28.04.2013 Views

Recovery Plan for the Mexican Spotted Owl (Strix occidentalis lucida )

Recovery Plan for the Mexican Spotted Owl (Strix occidentalis lucida )

Recovery Plan for the Mexican Spotted Owl (Strix occidentalis lucida )

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Recovery</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>Plan</strong><br />

<strong>for</strong> <strong>for</strong> <strong>the</strong><br />

<strong>the</strong><br />

<strong>Mexican</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong><br />

<strong>Owl</strong><br />

(<strong>Strix</strong> <strong>Strix</strong> <strong>occidentalis</strong> <strong>occidentalis</strong> <strong>lucida</strong> <strong>lucida</strong>) <strong>lucida</strong><br />

<strong>Plan</strong> de Recuperacion<br />

del Tecolate Moteado <strong>Mexican</strong>o<br />

(<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>)<br />

December 1995


Primary Primary Primary Authors:<br />

Authors:<br />

<strong>Recovery</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>Plan</strong><br />

<strong>for</strong> <strong>for</strong> <strong>the</strong><br />

<strong>the</strong><br />

<strong>Mexican</strong> <strong>Mexican</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong><br />

<strong>Owl</strong><br />

(<strong>Strix</strong> <strong>Strix</strong> <strong>occidentalis</strong> <strong>occidentalis</strong> <strong>lucida</strong> <strong>lucida</strong>) <strong>lucida</strong><br />

<strong>Plan</strong> de Recuperacion<br />

del Tecolate Moteado <strong>Mexican</strong>o<br />

(<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>)<br />

Nancy M. Kaufman, Regional Director,<br />

U.S. Department of <strong>the</strong> Interior, Fish and Wildlife Service,<br />

Southwestern Region<br />

William M. Block; Fernando Clemente; Jack F. Cully; James L. Dick, Jr.; Alan B. Franklin;<br />

Joseph L. Ganey; Frank P. Howe; W.H. Moir; Steven L. Spangle; Sarah E. Rinkevich; Dean L. Urban;<br />

Robert Vahle; James P. Ward, Jr.; and Gary C. White<br />

O<strong>the</strong>r O<strong>the</strong>r Contributors:<br />

Contributors:<br />

Kevin J. Cook (Technical Editor): Brian Geils (Disturbance Analyses);<br />

Kate W. Grandison (Meeting Facilitator); Tim Keitt (Landscape Analyses); Juan F. Martinez-Montoya<br />

(<strong>Mexican</strong> <strong>Recovery</strong> Units); Art Needleman and Jill Simmons (Layout and Design);<br />

Joyce V. Patterson (Illustrator); Tom Spalding (State of Arizona); Rich Teck (Forest Modeling);<br />

Steven Thompson (Southwestern Tribal Liaison); Brenda E. Witsell (Administrative Assistant).<br />

1995<br />

Approved:__________________________________________________ Date:_____________<br />

Regional Director, U.S. Fish and Wildlife Service


DISCLAIMER:<br />

DISCLAIMER:<br />

This <strong>Recovery</strong> <strong>Plan</strong> is not intended to provide details on all aspects of <strong>Mexican</strong> spotted owl management.<br />

The <strong>Recovery</strong> <strong>Plan</strong> outlines steps necessary to bring about recovery of <strong>the</strong> species. The <strong>Recovery</strong> <strong>Plan</strong> is not<br />

a “decision document” as defined by <strong>the</strong> National Environmental Policy Act (NEPA). It does not allocate<br />

resources on public lands. The implementation of <strong>the</strong> recovery plan is <strong>the</strong> responsibility of Federal and State<br />

management agencies in areas where <strong>the</strong> species occurs. Implementation is done through incorporation of<br />

appropriate portions of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> in agency decision documents such as <strong>for</strong>est plans, park management<br />

plans, and State game management plans. Such documents are <strong>the</strong>n subject to <strong>the</strong> NEPA process <strong>for</strong><br />

public review and selection of alternatives.<br />

LITERATURE LITERATURE CITATIONS:<br />

CITATIONS:<br />

This document should be referenced in literature citations as follows:<br />

USDI Fish and Wildlife Service. 1995. <strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl: Vol.I.<br />

Albuquerque, New Mexico. XXXpp.<br />

A fee <strong>for</strong> additional copies of this document will be charged depending on <strong>the</strong> number of pages<br />

and postage. Additional copies of this document may be purchased from:<br />

Fish and Wildlife Reference Service<br />

5430 Grosvenor Lane, Suite 110<br />

Be<strong>the</strong>sda, MD 20814<br />

(303) 492-6403<br />

1-800-582-3421


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

MEXICAN MEXICAN SPO SPOTTED SPO TED O OOWL<br />

O WL RECO RECOVER RECO VER VERY VER Y PL PLAN PL AN<br />

Table Table of of Contents<br />

Contents<br />

List List of of Tables ables ables ................................................................................................................................ ................................................................................................................................ vii<br />

vii<br />

List List of of F FFigur<br />

F igur igures igur es ............................................................................................................................... ............................................................................................................................... viii<br />

viii<br />

Executiv ecutiv ecutive ecutiv e S SSummar<br />

S ummar ummary ummar ....................................................................................................................... ....................................................................................................................... ix<br />

ix<br />

Ackno ckno cknowledgments<br />

ckno wledgments wledgments........................................................................................................................<br />

wledgments ........................................................................................................................ xiii<br />

xiii<br />

PAR AR ART AR T I: I: PL PLAN PL AN DE DEVEL DE VEL VELOPMENT<br />

VEL OPMENT<br />

A. A. RECO RECOVER RECO RECO VER VERY VERY<br />

Y PL PLANNING<br />

PL ANNING ................................................................................................ ................................................................................................ 1<br />

B. B. LISTING LISTING ............................................................................................................................ ............................................................................................................................ 2<br />

The Present or Threatened Destruction, Modificiation, or Curtailment<br />

of its Habitat or Range .....................................................................................................2<br />

Overutilization <strong>for</strong> Commercial, Recreational, Scientific, or Educational<br />

Purposes ..........................................................................................................................3<br />

Disease or Predation ........................................................................................................ 3<br />

Inadequacy of Existing Regulatory Mechanisms ...................................................................3<br />

O<strong>the</strong>r Natural or Manmade Factors Affecting Its Continued Existence ................................3<br />

C. C. PP<br />

PAST PP<br />

AST AND AND CURRENT CURRENT MANA MANAGEMENT MANA GEMENT OF OF THE<br />

THE<br />

MEXICAN MEXICAN SPO SPOTTED SPO TED O OOWL<br />

O OWL<br />

WL .......................................................................................... .......................................................................................... 4<br />

Fish and Wildlife Service ...................................................................................................... 4<br />

Forest Service ....................................................................................................................... 4<br />

Forest Service Southwestern Region (Region 3) ................................................................ 4<br />

Forest Service Rocky Mountain Region (Region 2) .......................................................... 5<br />

Forest Service Intermountain Region (Region 4) .............................................................. 5<br />

O<strong>the</strong>r Federal Agencies ........................................................................................................ 6<br />

National Park Service ....................................................................................................... 6<br />

Bureau of Land Management ........................................................................................... 6<br />

Department of Defense ................................................................................................... 6<br />

States ................................................................................................................................... 7<br />

Arizona............................................................................................................................ 7<br />

New Mexico .................................................................................................................... 7<br />

Colorado ......................................................................................................................... 7<br />

Utah ................................................................................................................................ 7<br />

Texas ............................................................................................................................... 8<br />

Tribes .................................................................................................................................. 8<br />

Mescalero Apache Tribe ................................................................................................... 8<br />

White Mountain Apache Tribe......................................................................................... 8<br />

San Carlos Apache Tribe .................................................................................................. 9<br />

Jicarilla Apache ................................................................................................................ 9<br />

Navajo Nation ................................................................................................................. 9<br />

Mexico ................................................................................................................................ 9<br />

D. D. CONSIDERA<br />

CONSIDERA<br />

CONSIDERATIONS CONSIDERA<br />

CONSIDERATIONS<br />

TIONS IN IN PL PLAN PL AN DE DEVEL DE DEVEL<br />

VEL VELOPMENT<br />

VEL OPMENT ....................................................... ....................................................... 11<br />

11<br />

<strong>Recovery</strong> Units .................................................................................................................... 11<br />

The Current Situation ......................................................................................................... 11<br />

<strong>Recovery</strong> <strong>Plan</strong> Duration ....................................................................................................... 12<br />

Conservation <strong>Plan</strong>s <strong>for</strong> O<strong>the</strong>r <strong>Spotted</strong> <strong>Owl</strong> Subspecies ......................................................... 13<br />

i


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Ecosystem Management ..................................................................................................... 15<br />

Economic Considerations ...................................................................................................16<br />

Human Intervention and Natural Processes ........................................................................17<br />

Differences Between <strong>the</strong> Draft and Final <strong>Recovery</strong> <strong>Plan</strong>s .....................................................17<br />

PAR AR ART AR T II: II: BIOL BIOLOGICAL BIOL OGICAL AND AND AND ECOL ECOLOGICAL ECOL OGICAL BA BACK BA CK CKGR CK GR GROUND GR OUND<br />

A. A. GENERAL GENERAL BIOL BIOLOGY BIOL OGY AND AND AND ECOL ECOLOGICAL ECOL OGICAL REL RELATIONSHIPS<br />

REL TIONSHIPS<br />

OF OF THE THE MEXICAN MEXICAN SPO SPOTTED SPO TED O OOWL<br />

O WL ......................................................................... ......................................................................... 19<br />

19<br />

Taxonomy ..........................................................................................................................19<br />

Description ........................................................................................................................21<br />

Distribution and Abundance ..............................................................................................21<br />

Habitat Associations ........................................................................................................... 26<br />

Nesting and Roosting Habitat.........................................................................................26<br />

Foraging Habitat ............................................................................................................27<br />

Territoriality and Home Range ...........................................................................................27<br />

Vocalizations ......................................................................................................................28<br />

Interspecific Competition ...................................................................................................28<br />

Feeding Habitats and Prey Ecology .....................................................................................28<br />

Reproductive Biology .........................................................................................................29<br />

Mortality Factors ................................................................................................................31<br />

Predation ........................................................................................................................31<br />

Starvation .......................................................................................................................31<br />

Accidents ........................................................................................................................31<br />

Disease and Parasites .......................................................................................................32<br />

Population Biology ............................................................................................................. 32<br />

Survival ..........................................................................................................................32<br />

Reproduction .................................................................................................................33<br />

Environmental Variation.................................................................................................33<br />

Population Trends ...........................................................................................................33<br />

Movements.........................................................................................................................33<br />

Seasonal Movements. ...................................................................................................... 33<br />

Natal Dispersal ...............................................................................................................33<br />

Landscape Pattern and Metapopulation Structure ...............................................................34<br />

Conclusions .......................................................................................................................34<br />

B. B. RECO RECOVER RECO VER VERY VER Y UNIT UNITS........................................................................................................<br />

UNIT UNIT ........................................................................................................ 36<br />

36<br />

United States ......................................................................................................................36<br />

Colorado Plateau ............................................................................................................36<br />

Sou<strong>the</strong>rn Rocky Mountains - Colorado...........................................................................40<br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico......................................................................40<br />

Upper Gila Mountains....................................................................................................44<br />

Basin and Range - West ..................................................................................................46<br />

Basin and Range - East ...................................................................................................46<br />

Mexico ...............................................................................................................................49<br />

Sierra Madre Occidental - Norte .....................................................................................49<br />

Sierra Madre Oriental- Norte .......................................................................................... 50<br />

Sierra Madre Occidental - Sur........................ .................................................................50<br />

Sierra Madre Oriental - Sur.............................................................................................51<br />

Eje Neovolcanico ............................................................................................................51<br />

ii


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

C. C. DEFINITIONS DEFINITIONS OF OF FOREST FOREST CO COVER CO VER TYP YP YPES YP ES ES.............................................................<br />

ES ............................................................. 52<br />

52<br />

Literature Review................................................................................................................52<br />

Ponderosa Pine ...............................................................................................................53<br />

Mixed-conifer .................................................................................................................53<br />

Spruce-fir .......................................................................................................................54<br />

O<strong>the</strong>r Forest Types of Interest .........................................................................................54<br />

Chihuahua Pine..........................................................................................................54<br />

Quaking Aspen ...........................................................................................................55<br />

Riparian Forests ..........................................................................................................55<br />

<strong>Plan</strong> Definitions .................................................................................................................55<br />

Ponderosa Pine Forest .....................................................................................................55<br />

Pine-oak Forest ...............................................................................................................55<br />

Mixed-conifer Forest.......................................................................................................56<br />

High-elevation Forests, Including Spruce-fir Forest .........................................................57<br />

Quaking Aspen...............................................................................................................57<br />

Key to Forest Types.............................................................................................................57<br />

D. D. CONCEPTU<br />

CONCEPTUAL CONCEPTU AL FRAME FRAMEWORK FRAME ORK FOR FOR RECO RECOVER RECO VER VERY....................................................<br />

VER .................................................... 59<br />

59<br />

Ecosystem or Landscape Management ................................................................................59<br />

Fire.................................................................................................................................60<br />

O<strong>the</strong>r Natural Disturbances ............................................................................................61<br />

Degradation of Riparian Forests ......................................................................................65<br />

Timber Harvest and Silvicultural Practices ..........................................................................65<br />

Historical Perspectives ....................................................................................................65<br />

Past Practices ..............................................................................................................65<br />

Forest <strong>Plan</strong>s ................................................................................................................66<br />

Habitat Trends ............................................................................................................66<br />

Data Availability .....................................................................................................66<br />

Trends in Forest Land Base and Timber Volume ...................................................... 67<br />

Trends in Forest Types ............................................................................................67<br />

Trends in Size-class Distibutions .............................................................................68<br />

Summary of Recent Habitat Trends ........................................................................68<br />

Silvicultural Practices and Forest Management ................................................................68<br />

Silviculture .................................................................................................................69<br />

Even-aged Management .......................................................................................... 69<br />

Uneven-aged Management......................................................................................70<br />

Development and Maintenance of Stratified Mixtures .............................................71<br />

Conclusions.................................................................................................................... 71<br />

Grazing .............................................................................................................................. 72<br />

Recreation .......................................................................................................................... 73<br />

Types of Recreation ........................................................................................................ 73<br />

Camping .................................................................................................................... 74<br />

Hiking ....................................................................................................................... 74<br />

Off-road Vehicles ....................................................................................................... 74<br />

Rock Climbing ........................................................................................................ . 74<br />

Wildlife Viewing and Photography........................................................................... . 74<br />

Recreation Summary .................................................................................................. .74<br />

Summary........................................................................................................................... .75<br />

iii


PAR AR ART AR T III: III: RECO RECOVER RECO VER VERY VER<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

A. A. DELISTING DELISTING .................................................................................................................... .................................................................................................................... 76 76<br />

76<br />

The Delisting Process ........................................................................................................ 76<br />

Delisting Criteria ............................................................................................................... 76<br />

Monitoring Population Trends ....................................................................................... 77<br />

O<strong>the</strong>r Considerations of Population Monitoring ........................................................ 79<br />

Monitoring Habitat Trends ............................................................................................ 79<br />

Long-term Management <strong>Plan</strong> ........................................................................................ 80<br />

Delisting at <strong>the</strong> <strong>Recovery</strong> Unit Level .............................................................................. 80<br />

Moderating and Regulating Threats. .......................................................................... 81<br />

Habitat Trends Within <strong>Recovery</strong> Units ....................................................................... 81<br />

B. B. GENERAL GENERAL APPR APPROACH<br />

APPR CH ................................................................................................ ................................................................................................ 82<br />

82<br />

Assumptions and Guiding Principles.................................................................................. 82<br />

Assumptions. ................................................................................................................. 83<br />

Guiding Principles ......................................................................................................... 84<br />

General Recommendations ................................................................................................ 84<br />

Protected Areas .............................................................................................................. 84<br />

Protected Activity Center (PAC) ................................................................................ 84<br />

Guidelines ............................................................................................................. 84<br />

Rationale ............................................................................................................... 89<br />

Steep Slopes (Outside of PACs) .................................................................................. 89<br />

Guidelines ............................................................................................................. 89<br />

Rationale ............................................................................................................... 89<br />

Reserved Lands .......................................................................................................... 90<br />

Guidelines ............................................................................................................. 90<br />

Rationale ............................................................................................................... 90<br />

Restricted Areas ............................................................................................................. 90<br />

Existing Conditions ................................................................................................... 91<br />

Reference Conditions ................................................................................................. 91<br />

Nesting and Roosting target/threshold conditions .................................................. 91<br />

Coarse Filter .............................................................................................................. 93<br />

Fine Filter .................................................................................................................. 94<br />

Overriding Guidelines ........................................................................................... 94<br />

Specific Guidelines ................................................................................................ 94<br />

Rationale ............................................................................................................... 95<br />

Riparian Communities............................................................................................... 95<br />

Guidelines ............................................................................................................. 95<br />

Rationale ............................................................................................................... 95<br />

O<strong>the</strong>r Forest and Woodland Types ................................................................................. 95<br />

Grazing Recommendations ................................................................................................ 96<br />

Grazing Guidelines ........................................................................................................ 96<br />

Rationale <strong>for</strong> Grazing Guidelines ................................................................................... 97<br />

Recreation Recommendations............................................................................................ 98<br />

Guidelines ..................................................................................................................... 98<br />

<strong>Recovery</strong> Unit Considerations ........................................................................................... 98<br />

Colorado Plateau ........................................................................................................... 98<br />

Potential Threats ........................................................................................................ 99<br />

Sou<strong>the</strong>rn Rocky Mountains - Colorado.......................................................................... 99<br />

iv


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Potential Threats ........................................................................................................ 99<br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico..................................................................... 99<br />

Potential Threats ....................................................................................................... 100<br />

Upper Gila Mountains.................................................................................................. 100<br />

Potential Threats ....................................................................................................... 100<br />

Basin and Range - West ................................................................................................ 101<br />

Potential Threats ....................................................................................................... 101<br />

Basin and Range - East ................................................................................................. 102<br />

Potential Threats ....................................................................................................... 102<br />

Mexico ......................................................................................................................... 103<br />

Potential Threats ....................................................................................................... 104<br />

Sierra Madre Occidental - Norte ........................................................................... 104<br />

Sierra Madre Oriental - Norte............................................................................... 104<br />

Sierra Madre Occidental - Sur............................................................................... 104<br />

Sierra Madre Oriental - Sur .................................................................................. 104<br />

Eje Neovolcanico .................................................................................................. 104<br />

C. C. MONIT MONITORING MONIT ORING PR PROCEDURES<br />

PR OCEDURES<br />

OCEDURES.................................................................................<br />

OCEDURES ................................................................................. 105<br />

105<br />

Habitat Monitoring .......................................................................................................... 105<br />

Macrohabitat ................................................................................................................ 105<br />

Microhabitat ................................................................................................................ 106<br />

Population Monitoring ..................................................................................................... 107<br />

Target Population ......................................................................................................... 107<br />

Sampling Units............................................................................................................. 107<br />

Sampling Procedures..................................................................................................... 108<br />

Stratification ............................................................................................................. 108<br />

Selection of Quadrats ................................................................................................ 108<br />

Sampling Within Quadrats ....................................................................................... 108<br />

Banding Birds ........................................................................................................... 109<br />

Statistical Analysis......................................................................................................... 109<br />

Estimation of Capture Probability ............................................................................. 109<br />

Estimation of Density Per Quadrat............................................................................ 110<br />

Estimation of Density Per Stratum ............................................................................ 110<br />

Variance Estimators ................................................................................................... 111<br />

Cormack-Jolly-Seber Models ..................................................................................... 111<br />

Personnel ...................................................................................................................... 111<br />

Training .................................................................................................................... 112<br />

Computerized Data Entry and Summarization.............................................................. 112<br />

Costs ............................................................................................................................ 113<br />

Potential Experiments ................................................................................................... 113<br />

Alternative Designs <strong>for</strong> Population Monitoring ............................................................. 114<br />

Drawing New Samples of Quadrats Each Year ........................................................... 114<br />

Conducting Surveys Less Often than Yearly ............................................................... 115<br />

Adaptive Sampling .................................................................................................... 115<br />

Conclusions ..................................................................................................................... 115<br />

D. D. A AACTIVIT<br />

A ACTIVIT<br />

CTIVIT CTIVITY-SP<br />

CTIVIT -SP -SPECIFIC -SPECIFIC<br />

ECIFIC RESEAR RESEARCH<br />

RESEAR CH ............................................................................ ............................................................................ 116<br />

116<br />

Role of <strong>the</strong> Scientific Process ............................................................................................ 116<br />

Important Ascpects of Experimental Design...................................................................... 117<br />

Limitations in Past Research on <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> ................................................ 118<br />

Research Needs................................................................................................................. 118<br />

Dispersal ...................................................................................................................... 119<br />

v


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Genetics. .......................................................................................................................... 120<br />

Habitat ............................................................................................................................. 120<br />

Population Biology ...........................................................................................................120<br />

Threats to <strong>Recovery</strong> .......................................................................................................... 120<br />

O<strong>the</strong>r Ecosystem Components ......................................................................................... 120<br />

E. E. SUMMAR SUMMARY SUMMAR SUMMAR Y OF OF RECO RECOVER RECO VER VERY VER ........................................................................................ ........................................................................................ 121<br />

121<br />

PAR AR ART AR T IV IV: IV : RECO RECOVER RECO RECO VER VERY VER Y PL PLAN PL AN IMPLEMENT<br />

IMPLEMENTATION<br />

IMPLEMENT TION TION<br />

A. A. IMPLEMENTING IMPLEMENTING L LLAWS,<br />

L WS, REGUL REGULATIONS, REGUL TIONS, AND AND A AAUTHORITIES<br />

A UTHORITIES ......................<br />

...................... ...................... 124<br />

124<br />

Endangered Species Act .................................................................................................... 124<br />

Section 4 ...................................................................................................................... 124<br />

Section 5 ...................................................................................................................... 125<br />

Section 6 ...................................................................................................................... 125<br />

Section 7 ...................................................................................................................... 125<br />

Section 8 ...................................................................................................................... 126<br />

Section 9 ...................................................................................................................... 126<br />

Section 10 .................................................................................................................... 126<br />

National Forest Management Act...................................................................................... 126<br />

National Environmental Policy Act ...................................................................................128<br />

Migratory Bird Treaty Act ................................................................................................. 128<br />

Tribal Lands ..................................................................................................................... 128<br />

State and Private Lands ..................................................................................................... 128<br />

Mexico ............................................................................................................................. 128<br />

B. B. IMPLEMENT<br />

IMPLEMENTATION IMPLEMENT<br />

IMPLEMENT TION O OOVERSIGHT<br />

O VERSIGHT ........................................................................... ........................................................................... 129<br />

129<br />

<strong>Recovery</strong> Unit Working Teams.......................................................................................... 129<br />

Continuing Duties of <strong>the</strong> <strong>Recovery</strong> Team ......................................................................... 120<br />

Centralized <strong>Spotted</strong> <strong>Owl</strong> In<strong>for</strong>mation Repository ............................................................. 130<br />

C. C. STEPDO STEPDOWN STEPDO STEPDO WN OUTLINE OUTLINE .............................................................................................. .............................................................................................. 131<br />

131<br />

D. D. D. IMPLEMENT<br />

IMPLEMENT<br />

IMPLEMENTATION IMPLEMENT TION AND AND COST COST SCHEDULE SCHEDULE ....................................................... ....................................................... 137<br />

137<br />

GL GLOSSAR GL OSSAR OSSARY OSSAR ..................................................................................................................... ..................................................................................................................... 145<br />

145<br />

LITERA LITERATURE LITERA TURE CITED CITED ................................................................................................... ................................................................................................... 152<br />

152<br />

APP APPENDIX APP ENDIX A: A: The The The R RReco<br />

R eco ecover eco er ery er y Team eam and and and Associates Associates ....................................................... ....................................................... 163 163<br />

163<br />

APP APPENDIX APP APPENDIX<br />

ENDIX B: B: Schedule Schedule of of MM<br />

Meetings MM<br />

eetings ............................................................................. ............................................................................. 165 165<br />

165<br />

APP APPENDIX APP ENDIX C: C: Schedule Schedule of of F FField<br />

F ield Visits Visits ......................................................................... ......................................................................... 166<br />

166<br />

APP APPENDIX APP ENDIX D: D: List List of of EE<br />

English E nglish and and Latin Latin N NNames<br />

N ames .......................................................... .......................................................... 167<br />

167<br />

APP APPENDIX APP ENDIX E: E: Agencies Agencies Agencies and and P PPersons<br />

P ersons Commenting Commenting on on D DDraft<br />

D raft R RReco<br />

RR<br />

eco ecover eco er ery ery<br />

y P P<strong>Plan</strong><br />

PP<br />

lan ............... ............... 170 170<br />

170<br />

vi


List List of of Tables<br />

Tables<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able ES.1 ES.1 Overview of management categories by vegetation type <strong>for</strong> lands not<br />

administratively reserved.................................................................................................................xi<br />

Table able I.D.1 I.D.1 Summary of relative <strong>Mexican</strong> spotted owl population size, timber harvest threat, and fire<br />

threat by <strong>Recovery</strong> Unit.................................................................................................................. 12<br />

Table able II.A.1. II.A.1. Historical records and minimum numbers of <strong>Mexican</strong> spotted owls<br />

found during planned surveys and incidental observations, reported by <strong>Recovery</strong> Unit<br />

and land ownership .................................................................................................................. 23-24<br />

Table able II.B.1. II.B.1. Land-ownership patterns (thousands of hectares) in <strong>Recovery</strong> Units<br />

within <strong>the</strong> United States .................................................................................................................41<br />

Table able II.B.2. II.B.2. Land-ownership patterns (thousands of hectares) in <strong>Recovery</strong> Units<br />

within Mexico ................................................................................................................................50<br />

Table able able II.D.1. II.D.1. Changes in <strong>the</strong> area (ha x 1,000 [ac x 1,000]) and distribution of<br />

<strong>for</strong>est types from <strong>the</strong> 1960s to 1980s on commercial <strong>for</strong>est lands within Arizona<br />

and New Mexico ............................................................................................................................67<br />

Table able II.D.2. II.D.2. Changes in <strong>the</strong> density (trees/ha [trees/ac]) and distribution of<br />

tree size classes from <strong>the</strong> 1960s to 1980s on commercial <strong>for</strong>est lands within Arizona<br />

and New Mexico ............................................................................................................................68<br />

Table able III.B.1 III.B.1 III.B.1 Target/threshold conditions <strong>for</strong> mixed-conifer and pine-oak <strong>for</strong>ests<br />

within restricted areas .....................................................................................................................92<br />

Table able IV IV.1 IV .1 Implementation and cost schedule ..................................................................... 139-144<br />

vii


List List List of of Figures<br />

Figures<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e II.A.1. II.A.1. Geographic range of <strong>the</strong> spotted owl ......................................................................20<br />

Figur igur igure igur e II.A.2. II.A.2. II.A.2. Current distribution of <strong>Mexican</strong> spotted owls in <strong>the</strong> United States<br />

based on planned surveys and incidental observations recorded from 1990 through 1993...............22<br />

Figur igur igure igur e II.A.3. II.A.3. Current distribution of <strong>Mexican</strong> spotted owls in Mexico based on<br />

planned surveys and incidental observations recorded from 1990 through 1993 .............................. 25<br />

Figur igur igure igur e II.A.4. II.A.4. Geographic variability in <strong>the</strong> food habits of <strong>Mexican</strong> spotted owls<br />

represented as relative frequencies of (a) (a) woodrats, (b) (b) voles, (c) (c) gophers, (d) (d) birds,<br />

(e) (e) bats, and (f (f) (f ) reptiles..................................................................................................................30<br />

Figur igur igure igur e II.B.1. II.B.1. II.B.1. <strong>Recovery</strong> Units within <strong>the</strong> United States.................................................................37<br />

Figur igur igure igur e II.B.2. II.B.2. <strong>Recovery</strong> Units within <strong>the</strong> Republic of Mexico ........................................................38<br />

Figur igur igure igur e II.B.3 II.B.3 Colorado Plateau <strong>Recovery</strong> Unit.............................................................................. 39<br />

Figur igur igur igure igur e II.B.4 II.B.4 Sou<strong>the</strong>rn Rocky Mountains - Colorado <strong>Recovery</strong> Unit ............................................42<br />

Figur igur igure igure<br />

e II.B.5 II.B.5 Sou<strong>the</strong>rn Rocky Mountains - New Mexico <strong>Recovery</strong> Unit .......................................43<br />

Figur igur igure igur e II.B.6 II.B.6 Upper Gila Mountains <strong>Recovery</strong> Unit .....................................................................45<br />

Figur igur igure igur e II.B.7 II.B.7 Basin and Range - West <strong>Recovery</strong> Unit ....................................................................47<br />

Figur igur igure igur e II.B.8 II.B.8 Basin and Range - East <strong>Recovery</strong> Unit .....................................................................48<br />

Figur igur igure igur e II.D.1 II.D.1 Historical record of number of total fires and number of<br />

natural fires. Data from USFS Southwestern Region ......................................................................62<br />

Figur igur igure igur e II.D.2 II.D.2 Historical record of fires over four hectares in size. Data from<br />

USFS Southwestern Region ............................................................................................................62<br />

Figur igur igur igure igur e II.D.3 II.D.3 Five-year running averages of area burned. Data from USFS<br />

Southwestern Region......................................................................................................................63<br />

Figur igur igure igur e III.A.1 III.A.1 Hypo<strong>the</strong>tical curve of <strong>the</strong> statistical power to detect a trend in a population ...........78<br />

Figur igur igure igur e III.B.1 III.B.1 Conceptualization of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> and needs <strong>for</strong> delisting<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl depicting <strong>the</strong> interdependency of population<br />

monitoring, habitat monitoring, and management recommendations .............................................83<br />

Figur igur igure igur e III.B.2 III.B.2 Generalization of protection stratgies by <strong>for</strong>est/vegetation type ...............................85<br />

Figur igur igure igur e III.B.3 III.B.3 Examples of protected activity center (PAC) boundaries from<br />

<strong>the</strong> Lincoln National Forest ............................................................................................................87<br />

viii


Volume 1/Executive Summary<br />

INTRODUCTION<br />

INTRODUCTION<br />

The <strong>Mexican</strong> spotted owl was listed as a<br />

threatened species on 15 April 1993. Two<br />

primary reasons were cited <strong>for</strong> <strong>the</strong> listing: historical<br />

alteration of its habitat as <strong>the</strong> result of<br />

timber management practices, specifically <strong>the</strong><br />

use of even-aged silviculture, plus <strong>the</strong> threat of<br />

<strong>the</strong>se practices continuing, as provided in National<br />

Forest <strong>Plan</strong>s. The danger of catastrophic<br />

wildfire was also cited as a potential threat <strong>for</strong><br />

additional habitat loss. Concomitant with <strong>the</strong><br />

listing of <strong>the</strong> <strong>Mexican</strong> spotted owl, a <strong>Recovery</strong><br />

Team was appointed by FWS Southwestern<br />

Regional Director John Rogers to develop a<br />

<strong>Recovery</strong> <strong>Plan</strong>. This report constitutes <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

This <strong>Recovery</strong> <strong>Plan</strong> provides a basis <strong>for</strong><br />

management actions to be undertaken by landmanagement<br />

agencies and Indian Tribes to<br />

remove recognized threats and recover <strong>the</strong><br />

spotted owl. Primary actions will be taken by <strong>the</strong><br />

USDA Forest Service, USDI Bureau of Land<br />

Management, USDI Fish and Wildlife Service,<br />

USDI Bureau of Indian Affairs, and sovereign<br />

American Indian Tribes. The Fish and Wildlife<br />

Service will oversee implementation of <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> through its authorities under <strong>the</strong><br />

Endangered Species Act.<br />

The Team made every ef<strong>for</strong>t to identify and<br />

consider all sources of in<strong>for</strong>mation in developing<br />

this plan. Previous plans developed <strong>for</strong> <strong>the</strong><br />

nor<strong>the</strong>rn spotted owl (Thomas et al. 1990, Bart<br />

et al. 1992) and <strong>the</strong> Cali<strong>for</strong>nia spotted owl<br />

(Verner et al. 1992) were considered in <strong>the</strong><br />

development of this <strong>Recovery</strong> <strong>Plan</strong>. The Team<br />

analyzed data that had not been evaluated<br />

previously and re-analyzed data when appropriate<br />

to ensure that in<strong>for</strong>mation was consistent or<br />

to address questions not considered in previous<br />

analyses of those data.<br />

RECOVERY RECOVERY RECOVERY GOAL<br />

GOAL<br />

The purpose of this <strong>Recovery</strong> <strong>Plan</strong> is to<br />

outline <strong>the</strong> steps necessary to remove <strong>the</strong> Mexi-<br />

EXECUTIVE EXECUTIVE SUMMARY<br />

SUMMARY<br />

ix<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

can spotted owl from <strong>the</strong> list of threatened<br />

species.<br />

THE THE RECOVERY RECOVERY PLAN<br />

PLAN<br />

The <strong>Recovery</strong> <strong>Plan</strong> contains five basic<br />

elements:<br />

1. A recovery goal and a set of delisting<br />

criteria that, when met, will allow <strong>the</strong><br />

<strong>Mexican</strong> spotted owl to be removed from<br />

<strong>the</strong> list of threatened species.<br />

2. Provision of three general strategies <strong>for</strong><br />

management that provide varying levels<br />

of habitat protection depending on <strong>the</strong><br />

owl’s needs and habitat use.<br />

3. Recommendations <strong>for</strong> population and<br />

habitat monitoring.<br />

4. A research program to address critical<br />

in<strong>for</strong>mation needs to better understand<br />

<strong>the</strong> biology of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

and <strong>the</strong> effects of anthropogenic activities<br />

on <strong>the</strong> owl and its habitat.<br />

5. Implementation procedures that<br />

specify oversight and coordination<br />

responsibilities.<br />

Each of <strong>the</strong>se elements is described briefly<br />

below.<br />

Delisting Delisting Criteria<br />

Criteria<br />

The primary threat to <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl leading to its listing as a threatened species<br />

was <strong>the</strong> alteration of its habitat in Arizona and<br />

New Mexico as <strong>the</strong> result of timber management,<br />

specifically even-aged management.<br />

<strong>Mexican</strong> spotted owls use a variety of habitats,<br />

but are typically associated with multi-canopied<br />

stands of mature mixed-conifer and ponderosa<br />

pine-Gambel oak <strong>for</strong>ests. Past logging using<br />

even-aged shelterwood prescriptions that included<br />

short rotations and <strong>the</strong> removal of large


volumes of timber greatly simplified stand<br />

structures which adversely affected >300,000 ha<br />

(800,000 ac) of spotted owl habitat (Fletcher<br />

1990). Existing Forest <strong>Plan</strong>s call <strong>for</strong> continued<br />

use of shelterwood harvests, potentially leading<br />

to continued loss of owl habitat. However, <strong>the</strong><br />

Team recognizes and is encouraged by recent<br />

ef<strong>for</strong>ts to amend existing Forest <strong>Plan</strong>s to deemphasize<br />

<strong>the</strong> use of even-aged silviculture and<br />

incorporate <strong>the</strong> management guidelines provided<br />

within this <strong>Recovery</strong> <strong>Plan</strong>.<br />

The range of <strong>the</strong> <strong>Mexican</strong> spotted owl was<br />

divided into six <strong>Recovery</strong> Units in <strong>the</strong> United<br />

States and five in Mexico. <strong>Recovery</strong> Units were<br />

based on various factors including biotic provinces,<br />

<strong>the</strong> spotted owl’s ecology, and management<br />

considerations. If delisting criteria are met,<br />

<strong>Recovery</strong> Units can be delisted separately. The<br />

following criteria must be met <strong>for</strong> delisting to be<br />

considered: (1) <strong>the</strong> population in <strong>the</strong> three most<br />

populated <strong>Recovery</strong> Units must be stable or<br />

increasing after 10 years of monitoring; (2)<br />

scientifically-valid habitat monitoring protocols<br />

are designed and implemented to assess (a) gross<br />

changes in habitat quantity across <strong>the</strong> range of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl, and (b) habitat modifications<br />

and habitat trajectories within treated<br />

stands; and (3) a long-term management plan is<br />

in place to ensure appropriate management <strong>for</strong><br />

<strong>the</strong> spotted owl and its habitat. If <strong>the</strong>se three<br />

criteria are met, <strong>the</strong>n <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

can be delisted within any <strong>Recovery</strong> Unit if<br />

threats have been moderated or regulated, and if<br />

habitat trends are stable or increasing.<br />

Levels Levels of of Protection<br />

Protection<br />

General recommendations are proposed <strong>for</strong><br />

three levels of management: protected areas,<br />

restricted areas, and o<strong>the</strong>r <strong>for</strong>est and woodland<br />

types (Table ES.1). Protected areas include a 243<br />

ha (600 ac) “Protected Activity Center” (PAC)<br />

placed at known or historical nest and/or roost<br />

sites, slopes >40% in mixed-conifer and pine-oak<br />

<strong>for</strong>ests that have not been harvested within <strong>the</strong><br />

past 20 years, and administratively reserved<br />

lands. Harvest of trees >22.4 cm (9 inches) dbh<br />

(diameter at breast height) is not allowed within<br />

protected areas, but light underburning is<br />

permitted on a case-specific basis as needed to<br />

Volume 1/Executive Summary<br />

x<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

reduce fuels. Also, a fire risk-abatement program<br />

is proposed to allow <strong>the</strong> treatments of fuels using<br />

a combination of fuel removal and fire. This<br />

management can be conducted initially within<br />

10% of <strong>the</strong> PACs, after which time <strong>the</strong> effectiveness<br />

of <strong>the</strong> program should be evaluated. Similar<br />

management can be conducted on steep slopes,<br />

but with no areal restrictions.<br />

Restricted areas include ponderosa pine-<br />

Gambel oak and mixed-conifer <strong>for</strong>ests and<br />

riparian environments. Target/threshold criteria<br />

are provided to define <strong>the</strong> proportion of <strong>the</strong><br />

landscape that should be in or approaching<br />

conditions suitable <strong>for</strong> nesting and roosting. The<br />

remainder of <strong>the</strong> landscape should be managed<br />

in such a way to allocate stands to ensure a<br />

sustained provision of nest and roost habitat<br />

through time. Broad guidelines <strong>for</strong> riparian<br />

systems emphasize <strong>the</strong> maintenance and restoration<br />

of riparian areas to ensure a mix of size and<br />

age classes.<br />

O<strong>the</strong>r <strong>for</strong>est and woodland types include<br />

ponderosa pine and spruce-fir <strong>for</strong>ests, pinyonjuniper<br />

woodlands, and aspen groves that are not<br />

included within PACs. No specific guidelines are<br />

proposed, but general recommendations are<br />

given to manage <strong>the</strong>se areas <strong>for</strong> landscape diversity<br />

within natural ranges of variation.<br />

Population Population and and Habitat Habitat Monitoring<br />

Monitoring<br />

The <strong>Recovery</strong> <strong>Plan</strong> provides a detailed<br />

program to monitor spotted owl populations<br />

and habitats. Both are key components of <strong>the</strong><br />

delisting criteria. Population monitoring is<br />

restricted to <strong>the</strong> three most populated <strong>Recovery</strong><br />

Units because <strong>the</strong>ir spotted owl populations<br />

meet sample size criteria <strong>for</strong> <strong>the</strong> monitoring<br />

design. Fur<strong>the</strong>r, <strong>the</strong>se <strong>Recovery</strong> Units comprise<br />

<strong>the</strong> core <strong>Mexican</strong> spotted owl population and<br />

<strong>the</strong> Team assumes that <strong>the</strong>ir population status<br />

reflects that of <strong>the</strong> entire population. The design<br />

presented in <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> entails <strong>the</strong> use of<br />

mark-recapture methodology on random quadrats<br />

to estimate key population parameters. The<br />

objectives of <strong>the</strong> habitat monitoring are (a) to<br />

track gross changes in habitat quality and quantity<br />

using remote sensing technology, and (b) to<br />

evaluate whe<strong>the</strong>r treatments meet <strong>the</strong> desired<br />

goal of setting stands on trajectories to become<br />

replacement habitat.


Table able ES.1.<br />

ES.1. Overview of management categories by vegetation type <strong>for</strong> lands not<br />

administratively reserved.<br />

Volume 1/Executive Summary<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Vegetation egetation In n N NNest<br />

N est S SSlope<br />

S lope Timber imber H HHar<br />

H ar arvest ar est est<br />

Within ithin PP<br />

Past PP<br />

ast Management anagement anagement<br />

Type ype Ar Area? Ar ea? 1 > 40% 40% 20 20 Years? ears? Categor ategor ategory ategor<br />

Any Yes Protected<br />

Mixed-conifer No Yes No Protected<br />

Pine-Oak No Yes No Protected<br />

Mixed-conifer No Yes Yes Restricted<br />

Pine-Oak No Yes Yes Restricted<br />

Mixed-conifer No No Restricted<br />

Pine-Oak No No Restricted<br />

Riparian No No Restricted<br />

Ponderosa Pine No No O<strong>the</strong>r types<br />

Spruce-Fir No No O<strong>the</strong>r types<br />

Pinyon-Juniper No No O<strong>the</strong>r types<br />

Aspen No No O<strong>the</strong>r types<br />

Oak No No O<strong>the</strong>r types<br />

1 Refers to land contained within a protected activity center.<br />

Activity-specific Activity-specific Research Research Program Program<br />

Program<br />

The <strong>Recovery</strong> Team made extensive use of<br />

scientific data. During <strong>the</strong> process of ga<strong>the</strong>ring<br />

and evaluating <strong>the</strong>se data, it became evident that<br />

additional in<strong>for</strong>mation was needed to refine <strong>the</strong><br />

recovery measures. Past research ef<strong>for</strong>ts emphasized<br />

inductive approaches to ga<strong>the</strong>r in<strong>for</strong>mation<br />

on basic life history needs of <strong>the</strong> spotted owl.<br />

Although <strong>the</strong>se research ef<strong>for</strong>ts provided some<br />

key in<strong>for</strong>mation, more rigorous and directed<br />

approaches will be needed to address questions<br />

on dispersal, genetics, habitat, populations, and<br />

effect of management on spotted owls and o<strong>the</strong>r<br />

ecosystem attributes.<br />

Implementation Implementation Measures<br />

Measures<br />

<strong>Recovery</strong> <strong>Plan</strong>s are not self-implementing<br />

under <strong>the</strong> Endangered Species Act. Thus, an<br />

implementation schedule is provided that<br />

outlines steps needed <strong>for</strong> <strong>the</strong> execution of <strong>the</strong><br />

recovery measures. These implementation<br />

guidelines include <strong>the</strong> <strong>for</strong>mation of an interagency<br />

working team <strong>for</strong> each <strong>Recovery</strong> Unit to<br />

oversee implementation of <strong>the</strong> recovery measures<br />

that encompass four broad areas: resource<br />

xi<br />

management programs, active management<br />

actions, monitoring, and research.<br />

CONCLUSION<br />

CONCLUSION<br />

The <strong>Recovery</strong> <strong>Plan</strong> is based largely on final<br />

and preliminary results of field studies of spotted<br />

owl habitat use, population biology, and distribution.<br />

The Team relied on in<strong>for</strong>mation published<br />

in <strong>the</strong> both scientific and “gray” literature.<br />

If data were available but unanalyzed, <strong>the</strong> Team<br />

made every reasonable ef<strong>for</strong>t to conduct those<br />

analyses. Reanalyses of data were conducted<br />

when <strong>the</strong> Team wished to address questions not<br />

addressed by those who collected <strong>the</strong> data. Thus,<br />

this <strong>Recovery</strong> <strong>Plan</strong> represents <strong>the</strong> current stateof-knowledge<br />

on <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

The <strong>Recovery</strong> <strong>Plan</strong> recommendations are a<br />

combination of (1) protection of both occupied<br />

habitats and unoccupied areas approaching<br />

characteristics of nesting habitat, and (2) implementation<br />

of ecosystem management within<br />

unoccupied but potential habitat. The goal is to<br />

protect conditions and structures used by spotted<br />

owls where <strong>the</strong>y exist and set o<strong>the</strong>r stands on<br />

a trajectory to grow into replacement nest<br />

habitat or to provide conditions <strong>for</strong> <strong>for</strong>aging and


dispersal. By necessity this <strong>Plan</strong> is a hybrid<br />

approach because <strong>the</strong> status of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl as a threatened species requires some<br />

level of protection until <strong>the</strong> subspecies is<br />

delisted. These constraints modify ways and<br />

opportunities to manage ecosystems within<br />

landscapes where owls occur or might occur in<br />

<strong>the</strong> future. The <strong>Plan</strong> advocates applying ecosystem<br />

management in two slightly different ways.<br />

Within unoccupied mixed-conifer and pine-oak<br />

<strong>for</strong>est on


ACKNOWLEDGMENTS<br />

ACKNOWLEDGMENTS<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Many people helped in <strong>the</strong> completion of this <strong>Plan</strong>. A list of people who commented on <strong>the</strong> draft<br />

<strong>Recovery</strong> <strong>Plan</strong> is provided in Appendix E. In addition, we thank everyone who contributed valuable<br />

time and in<strong>for</strong>mation to <strong>the</strong> <strong>Recovery</strong> Team, and apologize <strong>for</strong> any omissions in our special thanks to<br />

<strong>the</strong> following:<br />

Suppor uppor upport uppor t of of <strong>the</strong> <strong>the</strong> <strong>the</strong> team team: team Nancy M. Kaufman, John G. Rogers, Jim Young, Susan MacMullin, Jamie<br />

Rappaport-Clark (Fish and Wildlife Service [FWS] Region 2); Denver Burns (Rocky Mountain Station<br />

[RMS], Colorado); Charles W. Cartwright, Jim Lloyd, John Kirkpatrick, Larry Henson (Forest Service<br />

[FS] Region 3).<br />

Providing viding viding data data to to <strong>the</strong> <strong>the</strong> team team: team Craig Allen, Mike Britten, Dick Braun, Larry Henderson, Clive<br />

Pinnock (National Park Service [NPS]); George Alter, Cheryl Caro<strong>the</strong>rs, Larry Cordova, Larry Cosper,<br />

Ka<strong>the</strong>rin Farr, Keith Fletcher, Hea<strong>the</strong>r Hollis, Patrick D. Jackson, Mike Man<strong>the</strong>i, Renee Galeano-<br />

Popp, Danney Salas, John Shafer, Dennis Watson (FS); Mike Bogan, Cindy Ramotnik (National<br />

Biological Service, [NBS]); Erik Brekke (Bureau of Land Management, [BLM]); Russell Duncan (SW<br />

Field Biologists); R.J. Gutiérrez, David Olson, Mark Seamans (Humboldt State Univ. [HSU]); Charles<br />

Johnson, Richard Reynolds (RMS); Susan Eggen-McIntosh (Sou<strong>the</strong>rn Station New Orleans); Pat<br />

Mehlhop (Natural Heritage Program, NM); Hildy Reiser (Holloman AFB, NM); Steve Speich (Dames<br />

and Moore Consultants); Roger Skaggs (Glenwood, NM); Luis Tarango (Colegio De Postgraduados,<br />

San Luis Potosi, Mexico); Jim Tress (SWCA Inc.); Dave Willey (N. Arizona Univ. [NAU]); Rick<br />

Winslow (Navajo Fish and Wildl. Dept.).<br />

Consultation Consultation and and and P PPresentations<br />

P esentations esentations: esentations Cristen Allen, Don Reimer (DR Systems); Bruce Anderson,<br />

Douglas A. Boyce, Paul Boucher, Cecelia Dargan, Don DeLorenzo, Jim Ellenwood, Leon Fager, Mary<br />

Lou Fairwea<strong>the</strong>r, Leon Fisher, Keith Fletcher, Hea<strong>the</strong>r Hollis, Marlin Johnson, Milo Larson, Bob<br />

Leaverton, Mike Leonard, Jim Lloyd, Mike Man<strong>the</strong>i, Chris Mehling, Terry Myers, Danney Salas, Steve<br />

Servis, George Sheppard, Tom Skinner, Josh Taiz, Rich Teck, Borys Tkacz, Jon Verner, Dennis Watson<br />

(FS); Bill Austin, Steve Chambers, George Devine, Jennifer Fowler-Propst, Marcos Gorreson, Sonja<br />

Jahrsdoerfer, Michele James, Sue Linner, Britta Muiznieks, Carol Torrez (FWS); George Barrowclough<br />

(Am. Museum of Natural Hist.); Ernie Beil, Sheridan Stone (Ft. Huachuca, AZ); Erik Brekke (BLM);<br />

Tina Carlson (Natural Heritage Program); Mario Cirett (Ecological Center, Mexico); Martha Coleman,<br />

Tanya Shenk (Colorado State Univ.); Gerald Craig (CO Division of Wildlife); Norris Dodd (AZ Game<br />

and Fish Dept.); Russell Duncan (SW Field Biologists); Jeff Feen, Tim Wilhite (San Carlos Apache);<br />

Eric Forsman (FS Pacific NW Res. Stn.); Brian Geils, John Lundquist, Richard Reynolds (RMS); R.J.<br />

Gutiérrez, Dave Olson, Mark Seamans (HSU); Steve Haglund (Bureau of Indian Affairs); Craig Hauke<br />

(NPS); Tod Hull, Rich VanDemark (Applied Ecosystem Management Inc.); Terry Johnson (independent<br />

researcher); Joe Jojola (White Mtn. Apache); Norman Jojola (Mescalero Apache); Cameron<br />

Martinez (Nor<strong>the</strong>rn Pueblos); Marcelo Marquez, Luis Tarango (Colegio De Postgraduados, San Luis<br />

Potosi, Mexico); John McTague (NAU); Armando Cortez Ortiz, Jose Luis Omelasde Anda (National<br />

Institute of Statistics and Geographic Info., Mexico); Peter Stacey (Univ. of Nevada); Alejandro<br />

Velazques (Silviculturist, Mexico); Jerry Verner (FS Pacific SW Res. Stn.); Bob Waltermire (FWS/<br />

NBS); Sartor O. Williams, III (NM Dept. Game & Fish); Kendall Young (New Mexico State. Univ.);<br />

Mary Price (Univ. Calif., Riverside); Carl Marti (Weber State College); Andy Carey (FS Pacific NW<br />

Res. Stn.).<br />

xiii


xiv<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Conducting Conducting FF<br />

Field FF<br />

ield ield Trips ips ips: ips Bruce Anderson, Don DeLorenzo, Hea<strong>the</strong>r Green, Dave Nelson,<br />

Danney Salas, Steve Servis (FS); Russell Duncan (SW Field Biologists); Joe Jojola (White Mtn.<br />

Apache); San Carlos Apache Tribe, White Mountain Apache Tribe (providing <strong>the</strong> team access to <strong>the</strong>ir<br />

lands); Steve Speich (Dames and Moore Consultants); Sheridan Stone (Ft. Huachuca, AZ); Luis<br />

Tarango (Colegio De Postgraduados, San Luis Potosi, Mexico).<br />

Ga<strong>the</strong>r a<strong>the</strong>r a<strong>the</strong>ring a<strong>the</strong>r ing and and and Analyzing Analyzing D DData<br />

D ata ata: ata Damon Brown, Charlotte Corbett, Laura Duncan, Jill Dwyer,<br />

Jeff Jenness, Rudy King, Sharon Masek, Mike Nelson, Janie Sandoval, Peter Scott, Wayne Shepperd,<br />

Doug Spaeth, Mike Stanghellini, Gerry Wildeman, Randy Wilson, Rick Winslow, Karen Yarnell<br />

(RMS); Martha Coleman, Kevin Kalin (CSU); Susan DeRosier, Earl Hill, Karen Holmstrom, Rich<br />

Teck (FS); Susan Eggen-McIntosh (FS Sou<strong>the</strong>rn Stn., New Orleans); Bob Waltermire, Barb White<br />

(FWS/BS); Dave Willey (NAU).<br />

Logistical Logistical S SSuppor<br />

S uppor upport: uppor Bekki King, Karen Montano (FWS); Kim Ross (RMS).


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Volume I


<strong>Recovery</strong> <strong>Plan</strong> Development<br />

Volume I, Part I


The USDI Fish and Wildlife Service (FWS)<br />

added <strong>the</strong> <strong>Mexican</strong> spotted owl to <strong>the</strong> List of<br />

Threatened and Endangered Wildlife (50 CFR<br />

17.11) as a threatened species, effective on 15<br />

April 1993. Section 4(f)(1) of <strong>the</strong> Endangered<br />

Species Act of 1973, as amended (Act) (16<br />

U.S.C. 1531), requires <strong>the</strong> Secretary of <strong>the</strong><br />

Interior (usually delegated to <strong>the</strong> Director of <strong>the</strong><br />

FWS) to “...develop and implement (recovery)<br />

plans <strong>for</strong> <strong>the</strong> conservation of endangered species<br />

and threatened species...unless he finds that such<br />

a plan will not promote <strong>the</strong> conservation of <strong>the</strong><br />

species.”<br />

To develop scientifically credible recovery<br />

plans <strong>for</strong> listed species, <strong>the</strong> FWS may appoint<br />

recovery teams comprised of scientists and<br />

resource specialists with expertise ei<strong>the</strong>r on <strong>the</strong><br />

species being considered or with o<strong>the</strong>r relevant<br />

expertise. In <strong>the</strong> case of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl, <strong>the</strong> FWS appointed <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong><br />

<strong>Owl</strong> <strong>Recovery</strong> Team (<strong>Recovery</strong> Team). A list of<br />

<strong>Recovery</strong> Team members and <strong>the</strong>ir areas of<br />

expertise can be found in Appendix A. A chronology<br />

of <strong>Recovery</strong> Team activities is provided in<br />

Appendices B and C.<br />

<strong>Recovery</strong> teams present recovery plans to <strong>the</strong><br />

FWS as <strong>the</strong>ir recommendation on <strong>the</strong> steps<br />

necessary to remove a species from <strong>the</strong> List of<br />

Threatened and Endangered Wildlife and <strong>Plan</strong>ts.<br />

Removal from <strong>the</strong> list, or “delisting,” means <strong>the</strong><br />

species is no longer in need of protection under<br />

<strong>the</strong> Act and is <strong>the</strong>re<strong>for</strong>e considered “recovered.”<br />

If deemed acceptable, <strong>the</strong> Director of <strong>the</strong> FWS<br />

Region assigned <strong>the</strong> lead <strong>for</strong> that species approves<br />

<strong>the</strong> plan.<br />

The FWS, pursuant to requirements under<br />

section 4(f)(4) of <strong>the</strong> Act, published a Notice of<br />

Availability of <strong>the</strong> Draft <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> in <strong>the</strong> Federal Register on March<br />

27, 1995 (60 FR 15787). In addition to this<br />

general solicitation <strong>for</strong> in<strong>for</strong>mation and public<br />

comment, <strong>the</strong> FWS sent copies of <strong>the</strong> draft<br />

<strong>Recovery</strong> <strong>Plan</strong> to numerous Federal and State<br />

agencies, Indian Tribes, county governments,<br />

environmental and industry groups, and o<strong>the</strong>rs<br />

who had expressed interest in <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Finally, specific professional organizations<br />

and individuals were asked to provide<br />

Volume I/Part I<br />

A. A. RECOVERY RECOVERY RECOVERY PLANNING<br />

PLANNING<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

peer review of ei<strong>the</strong>r <strong>the</strong> entire document or<br />

portions treating subjects within <strong>the</strong>ir specific<br />

areas of expertise. A list of reviewers is provided<br />

in Appendix E.<br />

<strong>Recovery</strong> plans are nei<strong>the</strong>r self-implementing<br />

nor legally binding. Ra<strong>the</strong>r, approved recovery<br />

plans effectively constitute FWS policy on that<br />

listed species or group of species, <strong>the</strong>reby guiding<br />

<strong>the</strong> Service in conducting various processes<br />

required under <strong>the</strong> Act, such as section 7 consultation,<br />

conservation planning under section 10,<br />

and o<strong>the</strong>r procedures. In most cases, recovery<br />

plans are followed by o<strong>the</strong>r Federal agencies in<br />

compliance with <strong>the</strong> mandate under sections<br />

2(c)(1) and 7(a)(1) of <strong>the</strong> Act to utilize <strong>the</strong>ir<br />

authorities in carrying out programs <strong>for</strong> <strong>the</strong><br />

conservation of endangered and threatened<br />

species. In addition, State and local governments<br />

usually follow <strong>the</strong> recommendations of recovery<br />

plans in <strong>the</strong>ir species conservation ef<strong>for</strong>ts.<br />

Section 4(f)(1)(B) of <strong>the</strong> Act specifies <strong>the</strong><br />

contents of a recovery plan:<br />

“(i) a description of such site-specific<br />

management actions as may be necessary<br />

to achieve <strong>the</strong> plan’s goal <strong>for</strong> <strong>the</strong><br />

conservation and survival of <strong>the</strong><br />

species” (III.B);<br />

“(ii) objective, measurable criteria which,<br />

when met, would result in a<br />

determination...that <strong>the</strong> species be<br />

removed from <strong>the</strong> list” (III.A);<br />

“(iii) estimates of <strong>the</strong> time required and <strong>the</strong><br />

cost to carry out those measures<br />

needed to achieve <strong>the</strong> plan’s goal and<br />

to achieve intermediate steps toward<br />

that goal” (IV.C and IV.D).


The <strong>Mexican</strong> spotted owl is one of three<br />

spotted owl subspecies (see II.A). Under section<br />

3 of <strong>the</strong> Act, <strong>the</strong> term “species” includes “...any<br />

subspecies of fish or wildlife...”. Although <strong>the</strong><br />

<strong>Mexican</strong> spotted owl is a subspecies, it is sometimes<br />

referred to as a “species” in this document<br />

when discussed in <strong>the</strong> context of <strong>the</strong> Act or o<strong>the</strong>r<br />

laws and regulations. An “endangered species” is<br />

defined under <strong>the</strong> Act as “...any species which is<br />

in danger of becoming extinct throughout all or<br />

a significant portion of its range....” A threatened<br />

species is one “...which is likely to become<br />

an endangered species in <strong>the</strong> <strong>for</strong>eseeable future<br />

throughout all or a significant portion of its<br />

range.” Section 4 (A)(1) of <strong>the</strong> Act lists five<br />

factors that can, ei<strong>the</strong>r singly or collectively,<br />

result in listing as endangered or threatened:<br />

“(A) <strong>the</strong> present or threatened destruction,<br />

modification, or curtailment of its<br />

habitat or range;<br />

(B) overutilization <strong>for</strong> commercial, recreational,<br />

scientific, or educational<br />

purposes;<br />

(C) disease or predation;<br />

(D) <strong>the</strong> inadequacy of existing regulatory<br />

mechanisms;<br />

(E) o<strong>the</strong>r natural or man-made factors<br />

affecting its continued existence.”<br />

The final rule listing <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl as a threatened species (final rule) (58 FR<br />

14248) provides a detailed discussion of <strong>the</strong><br />

primary factors (A and D) leading to <strong>the</strong> determination<br />

of threatened status. It should be noted<br />

that <strong>the</strong> <strong>Recovery</strong> Team summarizes <strong>the</strong> final<br />

rule here <strong>for</strong> in<strong>for</strong>mation purposes only. The<br />

<strong>Recovery</strong> Team’s assessment of <strong>the</strong> current<br />

situation with regard to <strong>the</strong> subspecies’ status<br />

and threats is reflected in Part III. The following<br />

briefly summarizes <strong>the</strong> factors leading to <strong>the</strong><br />

species’ listing, as discussed in <strong>the</strong> final rule:<br />

Volume I/Part I<br />

B. B. LISTING<br />

LISTING<br />

2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

THE THE PRESENT PRESENT OR OR THREATENED<br />

THREATENED<br />

THREATENED<br />

DESTRUCTION, DESTRUCTION, MODIFICATION,<br />

MODIFICATION,<br />

OR OR CURTAILMENT CURTAILMENT CURTAILMENT OF OF ITS<br />

ITS<br />

HABITAT HABITAT OR OR OR RANGE RANGE<br />

RANGE<br />

Past, current, and future timber-harvest<br />

practices in <strong>the</strong> Southwestern Region (Region 3)<br />

of <strong>the</strong> USDA Forest Service (FS) were cited as<br />

<strong>the</strong> primary factors leading to listing <strong>the</strong> <strong>Mexican</strong><br />

spotted owl as a threatened species. The final<br />

rule stated that <strong>the</strong> Southwestern Region of <strong>the</strong><br />

FS managed timber primarily under a<br />

shelterwood harvest regime. This harvest method<br />

produces even-aged stands ra<strong>the</strong>r than <strong>the</strong><br />

uneven-aged, multi-layered stands most often<br />

used by <strong>Mexican</strong> spotted owls <strong>for</strong> nesting and<br />

roosting. In addition, <strong>the</strong> shelterwood silvicultural<br />

system calls <strong>for</strong> even-aged conditions in<br />

perpetuity. Thus, stands already changed from<br />

“suitable” to “capable” would not be allowed to<br />

return to a “suitable” condition; and acreage<br />

slated <strong>for</strong> future harvest will be similarly rendered<br />

perpetually unsuitable <strong>for</strong> <strong>Mexican</strong> spotted<br />

owl nesting and roosting.<br />

The final rule stated that “...significant<br />

portions of spotted owl habitat have been lost or<br />

modified,” and cited Fletcher (1990) in estimating<br />

that 420,000 ha (1,037,000 ac) of habitat<br />

were converted from “suitable” to “capable.” Of<br />

this, about 78.7%, or 330,000 ha (816,000 ac),<br />

was a result of human activities, whereas <strong>the</strong><br />

remainder was converted naturally, primarily by<br />

wildfire. According to <strong>the</strong> final rule, <strong>for</strong>est plans<br />

in <strong>the</strong> FS Region 3 allowed <strong>for</strong> up to 95% of<br />

commercial <strong>for</strong>est (59% of suitable spotted owl<br />

habitat) to be managed under a shelterwood<br />

system. The loss of lower- and middle-level<br />

riparian habitat plus habitat lost to recreation<br />

developments were also cited in <strong>the</strong> final rule as<br />

factors in habitat loss.


OVERUTILIZATION OVERUTILIZATION FOR<br />

FOR<br />

COMMERCIAL, COMMERCIAL, RECREATIONAL,<br />

RECREATIONAL,<br />

RECREATIONAL,<br />

SCIENTIFIC, SCIENTIFIC, OR OR EDUCATIONAL<br />

EDUCATIONAL<br />

PURPOSES<br />

PURPOSES<br />

The final rule stated that scientific research<br />

has <strong>the</strong> greatest potential <strong>for</strong> overutilization of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl, whereas birding,<br />

educational field trips, and agency “show me”<br />

trips are likely to increase as <strong>the</strong> owl becomes<br />

better known. The effects of <strong>the</strong>se activities,<br />

ei<strong>the</strong>r chronically or acutely, are unknown.<br />

Volume I/Part I<br />

DISEASE DISEASE OR OR PREDATION<br />

PREDATION<br />

The final rule stated that great horned owls<br />

and o<strong>the</strong>r raptors are predators of <strong>Mexican</strong><br />

spotted owls. It also implied that <strong>for</strong>est management<br />

created ecotones favored by great horned<br />

owls, thus creating an increased likelihood of<br />

contact between <strong>the</strong> two species.<br />

3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

INADEQUACY INADEQUACY OF OF EXISTING<br />

EXISTING<br />

REGULATORY REGULATORY MECHANISMS<br />

MECHANISMS<br />

MECHANISMS<br />

The final rule discussed various Federal and<br />

State laws and agency management policies,<br />

concluding that existing regulatory mechanisms<br />

were inadequate to protect <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl. For fur<strong>the</strong>r discussion on extant regulatory<br />

mechanisms, refer to Part IV.<br />

OTHER OTHER NATURAL NATURAL OR OR MANMADE<br />

MANMADE<br />

FACTORS FACTORS AFFECTING AFFECTING ITS<br />

ITS<br />

CONTINUED CONTINUED EXISTENCE<br />

EXISTENCE<br />

The final rule cited wildfires as a past and<br />

future threat to spotted owl habitat. The potential<br />

<strong>for</strong> increasing malicious and accidental<br />

anthropogenic harm to <strong>the</strong> species was also cited<br />

as a possible threat. In addition, <strong>the</strong> final rule<br />

recognized <strong>the</strong> potential <strong>for</strong> <strong>the</strong> barred owl to<br />

expand its range into that of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl, resulting in possible competition and/or<br />

hybridization. It was speculated that habitat<br />

fragmentation may encourage and hasten this<br />

expansion.


Volume I/Part I<br />

C. C. PAST PAST AND AND CURRENT CURRENT MANAGEMENT<br />

MANAGEMENT<br />

OF OF THE THE MEXICAN MEXICAN SPOTTED SPOTTED OWL<br />

OWL<br />

Prior to <strong>the</strong> proposed listing of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, some Federal agencies and involved<br />

States had conferred special status on <strong>the</strong> subspecies<br />

(e.g., State “threatened,” FS “sensitive,”<br />

FWS “candidate”) in recognition of its rarity,<br />

habitat preferences and threats to those habitat<br />

types, and/or need of special management<br />

considerations. This section summarizes <strong>the</strong><br />

special status assigned to <strong>the</strong> subspecies and <strong>the</strong><br />

resulting conservation ef<strong>for</strong>ts.<br />

FISH FISH AND AND WILDLIFE WILDLIFE SERVICE<br />

SERVICE<br />

The FWS listed <strong>the</strong> entire spotted owl<br />

species as a Category-2 candidate in its 6 January<br />

1989 Notice of Review (54 FR 554). Category-2<br />

candidates are those species that <strong>the</strong> FWS<br />

believes may qualify <strong>for</strong> listing as threatened or<br />

endangered but <strong>for</strong> which insufficient in<strong>for</strong>mation<br />

is available to support <strong>the</strong> required rulemaking<br />

process. The nor<strong>the</strong>rn subspecies was<br />

listed as threatened in 1990; <strong>the</strong> Cali<strong>for</strong>nia and<br />

<strong>Mexican</strong> subspecies remained in Category-2<br />

candidate status.<br />

The <strong>Mexican</strong> spotted owl was proposed <strong>for</strong><br />

listing as a threatened species on 4 November<br />

1991 (56 FR 56344) as a result of a status review<br />

prompted by a petition to list <strong>the</strong> subspecies.<br />

Following publication of <strong>the</strong> listing proposal, <strong>the</strong><br />

FWS attempted to develop a conservation<br />

agreement with involved Federal agencies to<br />

conserve <strong>the</strong> <strong>Mexican</strong> spotted owl. This ef<strong>for</strong>t<br />

was unsuccessful, so <strong>the</strong> final rule was published<br />

on 16 March 1993 (58 CFR 14248). Critical<br />

habitat was not determinable at <strong>the</strong> time of<br />

listing.<br />

Since <strong>the</strong> listing of <strong>the</strong> subspecies, <strong>the</strong> FWS<br />

has been conducting <strong>the</strong> processes associated<br />

with listed species under <strong>the</strong> Act, such as section<br />

7 consultation on Federal actions that may affect<br />

<strong>the</strong> subspecies, issuance of research permits<br />

under Section 10, and recovery planning under<br />

section 4, including funding of several research<br />

projects.<br />

Two petitions to delist <strong>the</strong> species have been<br />

reviewed by <strong>the</strong> FWS. In both cases, delisting<br />

was determined to be “not warranted” because<br />

4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>the</strong> petitions failed to present substantial scientific<br />

and commercial in<strong>for</strong>mation to support<br />

<strong>the</strong>ir assertion that <strong>the</strong> species should be<br />

delisted. Notices of those findings, including<br />

discussions of <strong>the</strong> issues raised in <strong>the</strong> petitions,<br />

were published in <strong>the</strong> Federal Register on 23<br />

September 1993 (58 FR 49467) and 1 April<br />

1994 (59 FR 15361).<br />

The FWS proposed critical habitat <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl on 7 December 1994 (59<br />

FR 63162), and published <strong>the</strong> final critical<br />

habitat rule on 6 June 1995 (60 FR 29914).<br />

Since that time, <strong>the</strong> FWS has been in consultation<br />

with action agencies on <strong>the</strong> effects of<br />

proposed and ongoing actions on critical habitat.<br />

FOREST FOREST FOREST SERVICE<br />

SERVICE<br />

The primary administrator of lands supporting<br />

<strong>Mexican</strong> spotted owls in <strong>the</strong> United States is<br />

<strong>the</strong> FS. Most spotted owls have been found<br />

within FS Region 3 (including 11 National<br />

Forests in Arizona and New Mexico). The Rocky<br />

Mountain (Region 2, including two National<br />

Forests in Colorado) and Intermountain (Region<br />

4, including three National Forests in Utah)<br />

Regions support fewer spotted owls.<br />

Forest Forest Service Service Southwestern Southwestern Region<br />

Region<br />

(Region (Region 3)<br />

3)<br />

Prior to <strong>the</strong> listing of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl, FS Region 3 issued detailed guidelines <strong>for</strong><br />

its management. Those guidelines were issued as<br />

Interim Directive Number 1 (ID No. 1) in June<br />

1989, <strong>the</strong>n revised and reissued as ID No. 2<br />

approximately one year later. Although ID No. 2<br />

expired in December 1991, FS Region 3 has<br />

continued managing under those guidelines.<br />

Interim Directive Number 2 guidelines<br />

required establishing management territories<br />

around all nesting and roosting spotted owls, as<br />

well as territorial owls detected at night <strong>for</strong><br />

which daytime locations were not recorded. All<br />

management territories except those on <strong>the</strong><br />

Lincoln and Gila National Forests had a 182-ha


(450 ac) core area surrounded by 627 ha (1,550<br />

ac) of <strong>the</strong> “best available” habitat, extending <strong>the</strong><br />

area to 809 ha (2,000 ac) per management<br />

territory. On <strong>the</strong> Lincoln and Gila National<br />

Forests, <strong>the</strong> 182-ha (450 ac) cores were augmented<br />

by an additional 425 ha (1,050 ac) of<br />

habitat, <strong>for</strong> a total management territory size of<br />

607 ha (1,500 ac).<br />

Except <strong>for</strong> road construction, habitat degradation<br />

was not allowed within management<br />

territory cores. In <strong>the</strong> remainder of <strong>the</strong> management<br />

territory management activities, including<br />

timber harvest, were limited to 209-314 ha<br />

(516-775 ac). The FS guidelines provided no<br />

protection <strong>for</strong> unoccupied habitat except in<br />

wilderness areas and administratively restricted<br />

lands.<br />

The FS Region 3 has been in <strong>the</strong> process of<br />

amending <strong>for</strong>est plans through <strong>the</strong> National<br />

Environmental Policy Act (NEPA) process to<br />

incorporate <strong>the</strong> management recommendations<br />

contained in this <strong>Recovery</strong> <strong>Plan</strong>. The <strong>Recovery</strong><br />

Team commends that ef<strong>for</strong>t.<br />

Forest Forest Forest Service Service Rocky Rocky Mountain Mountain Region<br />

Region<br />

(Region (Region (Region 2)<br />

2)<br />

Region 2 of <strong>the</strong> FS <strong>for</strong>med a task <strong>for</strong>ce in<br />

1992 to begin developing management guidelines<br />

<strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl. These management<br />

guidelines are still in draft <strong>for</strong>m and<br />

have not been <strong>for</strong>mally approved and adopted by<br />

FS Region 2. However, management activities<br />

continue to be examined on a case-by-case basis,<br />

and ID No. 2 may be used as a general guideline.<br />

Forest Forest Service Service Service Intermountain Intermountain Region<br />

Region<br />

(Region (Region (Region 4)<br />

4)<br />

Prior to <strong>the</strong> listing of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl, biologists from FS Region 4 and Utah’s<br />

o<strong>the</strong>r land and wildlife management agencies,<br />

plus owl researchers, <strong>for</strong>med <strong>the</strong> Utah <strong>Mexican</strong><br />

<strong>Spotted</strong> <strong>Owl</strong> Working Group (Working Group).<br />

The Working Group meets annually to identify<br />

and address issues pertaining to <strong>the</strong> management<br />

and conservation of <strong>Mexican</strong> spotted owls in<br />

Utah (Kate Grandison, FS, Cedar City, UT, pers.<br />

comm.). The Utah <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong><br />

Technical Team (Technical Team) was <strong>for</strong>med by<br />

Volume I/Part I<br />

5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>the</strong> Working Group to focus on spotted owl<br />

issues such as (1) potential impacts to <strong>Mexican</strong><br />

spotted owls in sou<strong>the</strong>rn Utah; (2) current<br />

research and future research needs and priorities;<br />

(3) inventory and monitoring protocols; (4)<br />

management suggestions suitable <strong>for</strong> application<br />

by all land management agencies in sou<strong>the</strong>rn<br />

Utah; and (5) dissemination of in<strong>for</strong>mation<br />

from <strong>the</strong> Working Group and Technical Team to<br />

management and administrative levels. The goals<br />

of <strong>the</strong> Technical Team, which is composed of<br />

biologists from <strong>the</strong> FWS, FS, USDI Bureau of<br />

Land Management (BLM), USDI National Park<br />

Service (NPS), Utah Division of Wildlife Resources<br />

(UDWR), and a researcher/technical<br />

consultant, are to provide land and wildlife<br />

managers with <strong>the</strong> in<strong>for</strong>mation necessary to<br />

ensure <strong>the</strong> protection of <strong>Mexican</strong> spotted owls<br />

and to suggest strategies <strong>for</strong> managing spotted<br />

owl habitat in Utah.<br />

The Technical Team developed “Suggestions<br />

<strong>for</strong> Management of <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> in<br />

Utah.” These suggestions were sent to line<br />

officers <strong>for</strong> approval on 5 August 1994. Management<br />

territories on <strong>the</strong> Manti-LaSal National<br />

Forest were established using <strong>the</strong>se suggestions,<br />

although ID No. 2 has also been adopted by<br />

Region 4. Interim Directive No. 2 was modified<br />

in March 1994 in Region 4 to change <strong>the</strong> survey<br />

protocol to include only potential breeding<br />

habitat in canyon areas below 2,590 m (8,500<br />

ft).<br />

According to <strong>the</strong> suggestions, management<br />

territory size should be 1,330 ha (3,350 ac) with<br />

355-ha (875 ac) core areas of canyon habitat. In<br />

addition, a 0.8-km (0.5 mi) protection area<br />

centered on <strong>the</strong> nest site was established to<br />

protect <strong>the</strong> nest stand and surrounding areas.<br />

Habitat degradation is not allowed in <strong>the</strong> management<br />

territory areas. A “potential dispersal<br />

area” extends 58 km (35.8 mi) beyond <strong>the</strong><br />

perimeter of <strong>the</strong> management territory. This area<br />

can be used <strong>for</strong> timber harvest, but post-harvest<br />

conditions must meet a reasonable facsimile of<br />

<strong>the</strong> 50-11-40 dispersal rule developed by Thomas<br />

et al. (1990). Forest Service Region 4<br />

continues to manage under ID No. 2, with <strong>the</strong><br />

above modifications, except where superseded by<br />

<strong>the</strong> “Suggestions <strong>for</strong> Management of <strong>Mexican</strong><br />

<strong>Spotted</strong> <strong>Owl</strong>s in Utah.”


OTHER OTHER OTHER FEDERAL FEDERAL AGENCIES<br />

AGENCIES<br />

National National Park Park Service<br />

Service<br />

Several NPS-administered units are known<br />

to support <strong>Mexican</strong> spotted owls, including<br />

Zion, Capitol Reef, and Canyonlands National<br />

Parks plus Glen Canyon National Recreation<br />

Area in Utah; Mesa Verde National Monument<br />

in Colorado; Grand Canyon National Park plus<br />

Saguaro, Walnut Canyon, and Chiricahua<br />

National Monuments in Arizona; Bandelier<br />

National Monument in New Mexico; and<br />

Guadalupe Mountains National Park in Texas.<br />

The National Park Service Organic Act<br />

protects all wildlife on National Parks and<br />

Monuments. However, no specific management<br />

guidelines are in place <strong>for</strong> <strong>Mexican</strong> spotted owls,<br />

and <strong>the</strong> effectiveness of applying general laws<br />

and policies <strong>for</strong> spotted owls is difficult to<br />

evaluate.<br />

Bureau Bureau of of Land Land Land Management<br />

Management<br />

The BLM has developed management<br />

policies specifically <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

in Colorado and New Mexico. The Colorado<br />

guidelines state that “...in areas with a confirmed<br />

nest or roost site, surface management activities<br />

will be limited and will be determined on a caseby-case<br />

basis to allow as much flexibility as<br />

possible outside of <strong>the</strong> core area.” The BLM in<br />

Colorado has management guidelines <strong>for</strong> oil and<br />

gas development where <strong>Mexican</strong> spotted owls are<br />

known to occur. No surface occupancy is allowed<br />

within 0.4 km (0.25 mi) of a nest or roost<br />

site, and restrictions on o<strong>the</strong>r associated activities<br />

apply between 1 February and 31 July. <strong>Spotted</strong><br />

owl management policy by <strong>the</strong> BLM in New<br />

Mexico establishes and preserves cores of habitat<br />

wherever <strong>the</strong> owl is found. The BLM determines<br />

<strong>the</strong> size of <strong>the</strong> cores on a case-by-case basis.<br />

The BLM in Colorado follows <strong>the</strong> survey<br />

techniques of <strong>the</strong> FS Region 3 spotted owl<br />

protocol. Management territories have not been<br />

designated <strong>for</strong> known birds. Surveys are conducted<br />

in areas of potential habitat where<br />

projects are planned that may be in conflict with<br />

spotted owl management.<br />

Volume I/Part I<br />

6<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The BLM in Utah has no specific internal<br />

guidelines on management practices <strong>for</strong> <strong>Mexican</strong><br />

spotted owls. However, agency personnel did<br />

participate in producing “Suggestions <strong>for</strong> <strong>the</strong><br />

Management of <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong>s in<br />

Utah.” The BLM will incorporate management<br />

prescriptions <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl and its<br />

potential habitat into resource management<br />

plans as <strong>the</strong>y are updated over <strong>the</strong> next several<br />

years.<br />

The BLM in Arizona has no specific guidelines<br />

<strong>for</strong> managing <strong>Mexican</strong> spotted owls.<br />

However, <strong>the</strong> standard BLM procedure <strong>for</strong><br />

assessing impacts on threatened or endangered<br />

species will be followed <strong>for</strong> projects proposed in<br />

spotted owl habitat. Guidelines <strong>for</strong> protecting<br />

<strong>the</strong> owl or its habitat would <strong>the</strong>n be developed<br />

on a site-specific basis (Ted Corderey, BLM,<br />

Endangered Species Coordinator, Phoenix<br />

Office, pers. comm.).<br />

Department Department of of Defense Defense<br />

Defense<br />

The Fort Huachuca Military Reservation<br />

(Post) in sou<strong>the</strong>astern Arizona is <strong>the</strong> only military<br />

land known to support nesting <strong>Mexican</strong><br />

spotted owls. On <strong>the</strong> Post, military activity in<br />

spotted owl habitat is generally confined to<br />

various foot maneuvers, although <strong>the</strong> Army is<br />

considering expanding some tank maneuvers<br />

into higher elevations where <strong>the</strong> owl occurs<br />

(Sheridan Stone, Fort Huachuca Military Reservation,<br />

pers. comm.). One spotted owl site has<br />

been popular with birders <strong>for</strong> a number of years,<br />

but <strong>the</strong> effect of this activity is unknown. The<br />

Army also considers wildfire to be a potential<br />

threat and assesses <strong>the</strong> possibility of wildfire<br />

ignition when designing military activities on<br />

<strong>the</strong> Post.<br />

Wintering <strong>Mexican</strong> spotted owls have been<br />

found on Fort Carson, near Colorado Springs,<br />

Colorado, and breeding owls are present on <strong>the</strong><br />

Fremont Military Operating Area, which includes<br />

FS and BLM lands designated <strong>for</strong> conducting<br />

military maneuvers. Finally, low-level<br />

military air operations in some areas have been<br />

identified as actions that may affect <strong>Mexican</strong><br />

spotted owls. Such operations are likely to<br />

increase in <strong>the</strong> next several years, and <strong>the</strong> Department<br />

of Defense is currently funding studies


of <strong>the</strong> effects of <strong>the</strong>se activities on spotted owls<br />

(M. Hildegard Reiser, Holoman Air Force Base,<br />

pers. comm.).<br />

Arizona<br />

Arizona<br />

Volume I/Part I<br />

STATES<br />

STATES<br />

The <strong>Mexican</strong> spotted owl is listed as “threatened”<br />

on <strong>the</strong> list of “Threatened Native Wildlife<br />

in Arizona” (Arizona Game and Fish Department<br />

1988). The Arizona Game and Fish<br />

Department has authority to manage wildlife<br />

under provisions of Arizona Revised Statute 17,<br />

<strong>the</strong> goal of which is to maintain State’s natural<br />

biotic diversity by listing and protecting threatened<br />

and endangered species. “Threatened”<br />

species is defined as “...those species or subspecies<br />

whose continued presence in Arizona could<br />

be in jeopardy in <strong>the</strong> near future. Serious threats<br />

have been identified and populations are (a)<br />

lower than <strong>the</strong>y were historically or (b) extremely<br />

local and small.”<br />

Threatened status provides no special protection<br />

to species, although it does provide a<br />

mechanism through which <strong>the</strong> state can allocate<br />

Heritage Program grants to fund research <strong>for</strong><br />

specially designated species. However, general<br />

Arizona wildlife rules make it unlawful “...unless<br />

o<strong>the</strong>rwise prescribed...<strong>for</strong> a person to...take,<br />

possess, transport, buy, sell or offer or expose <strong>for</strong><br />

sale wildlife, except as expressly permitted ....”<br />

New New Mexico<br />

Mexico<br />

The State of New Mexico confers no special<br />

status on spotted owls. However, New Mexico<br />

Statute 17-2-14 makes it unlawful “...<strong>for</strong> any<br />

person to take, possess, trap or ensnare, or in any<br />

manner to injure, maim or destroy birds of <strong>the</strong><br />

order Strigi<strong>for</strong>mes.” However, permits may be<br />

obtained to take owls <strong>for</strong> purposes of Indian<br />

religion, scientific study, or falconry. In addition,<br />

persons who commercially raise poultry or game<br />

birds may legally kill any owl that has killed <strong>the</strong>ir<br />

stock.<br />

7<br />

Colorado Colorado<br />

Colorado<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The <strong>Mexican</strong> spotted owl was listed as<br />

threatened by <strong>the</strong> Colorado Division of Wildlife<br />

(CDOW) in 1993. “Threatened” wildlife is<br />

defined as “...any species or subspecies of wildlife<br />

which, as determined by <strong>the</strong> Colorado Wildlife<br />

Commission, is not in immediate jeopardy of<br />

extinction but is vulnerable because it exists in<br />

such small numbers or is so extremely restricted<br />

throughout all or a significant portion of its<br />

range that it may become endangered.” Threatened<br />

status protects wildlife species by making<br />

it unlawful “...<strong>for</strong> any person to take, possess,<br />

transport, export, process, sell or offer <strong>for</strong><br />

sale...any species or subspecies of [threatened]<br />

wildlife....” In addition, <strong>the</strong> CDOW is legislatively<br />

mandated to “...establish such programs<br />

including acquisition of land...as are deemed<br />

necessary <strong>for</strong> management of...threatened<br />

species.” An interagency working group<br />

coordinates spotted owl inventories throughout<br />

Colorado.<br />

Utah Utah<br />

Utah<br />

The UDWR included <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl as a sensitive species on its 1987 Native Utah<br />

Wildlife Species of Special Concern list (UDWR<br />

1987). “Sensitive” wildlife is defined as “...any<br />

wildlife species which, although still occurring in<br />

numbers adequate <strong>for</strong> survival, whose population<br />

has been greatly depleted, is declining in numbers,<br />

distribution, and/or habitat (S1); or occurs<br />

in limited areas and/or numbers due to a restricted<br />

or specialized habitat (S2).” A management<br />

program, including protection or enhancement,<br />

is needed <strong>for</strong> <strong>the</strong>se sensitive species.<br />

The owl’s status was elevated to “Threatened”<br />

in <strong>the</strong> revised draft list in 1992 (UDWR<br />

1992). According to UDWR definition, “threatened”<br />

species include “...any wildlife species,<br />

subspecies, or population which is likely to<br />

become an endangered species within <strong>the</strong> <strong>for</strong>eseeable<br />

future throughout all or a significant<br />

portion of its range in Utah or <strong>the</strong> world.”<br />

Both sensitive and threatened species receive<br />

“protected” status under Utah’s wildlife codes.<br />

For species under protected status, “...[A] person


may not take...protected wildlife or <strong>the</strong>ir parts;<br />

an occupied nest of protected wildlife; or an<br />

egg of protected wildlife.” Nor may a person<br />

“...transport,...sell or purchase...or possess<br />

protected wildlife or <strong>the</strong>ir parts.”<br />

Texas<br />

Texas<br />

Few <strong>Mexican</strong> spotted owls are documented<br />

<strong>for</strong> Texas, and most of <strong>the</strong> location records are in<br />

Guadalupe National Park. Thus, <strong>the</strong> State of<br />

Texas has no spotted owl program. However,<br />

Texas Parks and Wildlife Code Section 64.002<br />

provides protection <strong>for</strong> nongame birds by<br />

prohibiting killing, trapping, transportation,<br />

possession of parts, and <strong>the</strong> like. Destruction of<br />

eggs and nests of nongame birds is also prohibited.<br />

Volume I/Part I<br />

TRIBES<br />

TRIBES<br />

Tribal beliefs and philosophies guide resource<br />

management on Tribal lands. Several<br />

Tribes consider owls a bad omen; however, Tribal<br />

beliefs also dictate that all living creatures are<br />

essential parts of nature and, as such, <strong>the</strong>y are<br />

revered and protected. For example, <strong>the</strong> Elders<br />

Council of San Carlos Apache Tribe expressed<br />

<strong>the</strong> traditional view that owls and <strong>the</strong>ir homes<br />

should not be disturbed.<br />

<strong>Mexican</strong> spotted owl habitat or potential<br />

habitat exists on 10 Indian reservations in <strong>the</strong><br />

Southwest. Eight of <strong>the</strong> Tribes have conducted<br />

spotted owl surveys, and five Tribes have located<br />

spotted owls on <strong>the</strong>ir lands. Two o<strong>the</strong>r Tribes<br />

have historical spotted owl records.<br />

Reservations were established <strong>for</strong> <strong>the</strong> benefit<br />

of <strong>the</strong> Tribes and <strong>the</strong>ir members. Tribal lands are<br />

held in “trust” by <strong>the</strong> Federal Government. They<br />

are not considered public lands or part of <strong>the</strong><br />

public domain. Tribes are sovereign governments<br />

with management authority over wildlife and<br />

o<strong>the</strong>r Tribal land resources. Many Tribes maintain<br />

professionally staffed wildlife and natural<br />

resources management programs to ensure<br />

prudent management and protection of tribal<br />

resources, including threatened and endangered<br />

species.<br />

8<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The FWS is aware of spotted owl conservation<br />

ef<strong>for</strong>ts on five Indian reservations: <strong>the</strong><br />

Mescalero Apache, Fort Apache, San Carlos<br />

Apache, Jicarilla Apache, and Navajo Nation.<br />

Mescalero Mescalero Mescalero Apache Apache Tribe<br />

Tribe<br />

The Mescalero Apache Tribe in New Mexico<br />

actively manages <strong>the</strong>ir <strong>for</strong>est while managing <strong>for</strong><br />

all federally listed or proposed threatened or<br />

endangered species that may exist on <strong>the</strong> reservation,<br />

including <strong>the</strong> <strong>Mexican</strong> spotted owl. This is<br />

accomplished through developing strategies <strong>for</strong><br />

identifying and managing habitat determined by<br />

<strong>the</strong> Tribe to be necessary to ensure protection.<br />

The Mescalero has been working with <strong>the</strong> FWS<br />

in development of a conservation strategy <strong>for</strong> <strong>the</strong><br />

subspecies on reservation lands.<br />

White White Mountain Mountain Mountain Apache Apache Tribe<br />

Tribe<br />

The Tribe recently developed a conservation<br />

plan <strong>for</strong> <strong>Mexican</strong> spotted owls on <strong>the</strong> reservation.<br />

Areas containing spotted owls are placed in<br />

one of two land-management categories, termed<br />

Designated Management Areas (DMAs). Areas<br />

supporting “clusters” of four or more territories<br />

are considered “Category-1” DMAs. In <strong>the</strong>se<br />

areas, spotted owl habitat concerns drive management<br />

prescriptions; timber harvest is a<br />

secondary objective. Category-1 DMAs range<br />

from about 2,430-4,050 ha (6,000-10,000 ac),<br />

and contain 57% of known spotted owl sites on<br />

<strong>the</strong> reservation.<br />

“Category-2” DMAs include areas supporting<br />

1-3 owl territories. Habitat outside <strong>the</strong><br />

territories is managed only secondarily <strong>for</strong><br />

spotted owls, with o<strong>the</strong>r resource objectives<br />

given priority. No timber harvest is allowed in<br />

30-ha (75 ac) patches around owl activity centers.<br />

A seasonal restriction on potentially disturbing<br />

activities is provided in a 202-ha (500 ac)<br />

area, and timber prescriptions within this area<br />

should be designed to improve habitat integrity.<br />

The Tribe continues to survey <strong>the</strong>ir lands <strong>for</strong><br />

spotted owls. If more owl sites are detected,<br />

Category-1 and Category-2 DMAs may be<br />

established upon approval by <strong>the</strong> Tribal Council.


San San Carlos Carlos Apache Apache Tribe<br />

Tribe<br />

<strong>Spotted</strong> owl surveys on <strong>the</strong> San Carlos<br />

Apache Reservation have been conducted according<br />

to <strong>the</strong> FS Region 3 <strong>Mexican</strong> <strong>Spotted</strong><br />

<strong>Owl</strong> Inventory Protocol. <strong>Mexican</strong> spotted owl<br />

habitat has been identified and delineated<br />

throughout <strong>the</strong> reservation. A joint Tribal/<br />

Bureau of Indian Affairs interdisciplinary team<br />

evaluates effects of actions on spotted owls. Any<br />

potential impact on spotted owls or owl habitat<br />

is deferred until compliance with <strong>the</strong> Act and<br />

associated regulations is attained.<br />

Preliminary discussions between <strong>the</strong> San<br />

Carlos Apache and <strong>the</strong> FWS have taken place<br />

regarding development of specific spotted owl<br />

management guidelines. Approximately 90% of<br />

tribally identified nesting, roosting, and <strong>for</strong>aging<br />

habitats are on lands inoperable <strong>for</strong> timber<br />

harvest and <strong>the</strong>re<strong>for</strong>e are not in <strong>the</strong> commercial<br />

timber base.<br />

Jicarilla Jicarilla Jicarilla Apache Apache Apache Tribe<br />

Tribe<br />

The Jicarilla Apache Tribe has developed a<br />

spotted owl conservation plan, approved by <strong>the</strong><br />

Jicarilla Tribal Council and accepted by <strong>the</strong><br />

FWS. No resident owls have been detected to<br />

date on <strong>the</strong> reservation; however, in <strong>the</strong> event<br />

resident owls are detected, <strong>the</strong> Tribe has proposed<br />

to designate a 405-ha (1,000 ac) management<br />

territory. Uneven-aged timber management<br />

will be allowed to continue in all but 40 ha<br />

(100 ac) of <strong>the</strong> territory. In <strong>the</strong> absence of<br />

confirmed resident owls, all mixed-conifer stands<br />

of 10 ha (25 ac) or greater are treated as roosting/nesting<br />

sites, and timber harvest will not be<br />

allowed. A seasonal restriction around any active<br />

nest sites that are found is also proposed.<br />

Navajo Navajo Nation<br />

Nation<br />

<strong>Mexican</strong> spotted owl management on <strong>the</strong><br />

Navajo Nation, and particularly on <strong>the</strong> Navajo<br />

Nation Commercial Forest, currently adheres to<br />

FS Region 3’s ID No. 2. The FS Region 3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> Inventory Protocol is<br />

followed <strong>for</strong> all timber sales on <strong>the</strong> commercial<br />

<strong>for</strong>est and <strong>for</strong> any project or disturbance, on or<br />

Volume I/Part I<br />

9<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

off <strong>the</strong> commercial <strong>for</strong>est, that may impact<br />

spotted owls. The current Navajo spotted owl<br />

inventory program is limited to areas where<br />

timber sales or o<strong>the</strong>r projects are planned.<br />

The Navajo Nation is developing a multispecies<br />

conservation plan, including management<br />

guidelines <strong>for</strong> spotted owl conservation, in<br />

conjunction with <strong>the</strong>ir 10-year plan <strong>for</strong> managing<br />

commercial <strong>for</strong>est. Upon completion of <strong>the</strong><br />

multi-species conservation plan, <strong>the</strong> Navajo<br />

Nation may apply to <strong>the</strong> FWS <strong>for</strong> a section<br />

10(a)(1)(B) permit, which will allow limited<br />

incidental take to occur provided an adequate<br />

habitat conservation plan is implemented.<br />

MEXICO MEXICO<br />

MEXICO<br />

The <strong>Mexican</strong> spotted owl is listed as a<br />

threatened species under Mexico’s Official<br />

<strong>Mexican</strong> Norm (NOM) (NOM-059-ECOL-<br />

1994). Threatened species are defined as those<br />

which could face danger of extinction if <strong>the</strong><br />

conditions that cause deterioration or modification<br />

of <strong>the</strong>ir habitats, or decline of <strong>the</strong>ir populations,<br />

prevail.<br />

Species listed under NOM are af<strong>for</strong>ded<br />

certain protections:<br />

1. Possession, use, or derivation of profit<br />

from live wildlife or plants, whe<strong>the</strong>r<br />

originating in captivity or in <strong>the</strong> wild, are<br />

prohibited.<br />

2. Use and exploitation of <strong>the</strong> habitats of<br />

listed species are prohibited in some<br />

States.<br />

Some use of threatened species is allowed <strong>for</strong><br />

scientific and recovery purposes. For example,<br />

specimens and <strong>the</strong>ir parts, products, and byproducts<br />

can be removed from <strong>the</strong>ir natural<br />

environment <strong>for</strong> scientific purposes under<br />

permits issued by legal authorities, with <strong>the</strong><br />

understanding that specimens or <strong>the</strong>ir parts<br />

cannot be used <strong>for</strong> commercial purposes. In<br />

addition, specimens can be removed from <strong>the</strong><br />

wild <strong>for</strong> <strong>the</strong> purpose of captive breeding upon<br />

approval of <strong>the</strong> <strong>Mexican</strong> government.<br />

Under NOM, recovery plans have been<br />

prepared <strong>for</strong> sea turtles and <strong>the</strong> monarch butter-


fly. The Government and o<strong>the</strong>r institutions have<br />

shown interest in <strong>the</strong> conservation of o<strong>the</strong>r<br />

species such as manatees, yellow-headed parrots,<br />

and <strong>Mexican</strong> spotted owls, but currently no<br />

recovery plans are in place <strong>for</strong> <strong>the</strong>se species.<br />

Social, economic, and political systems differ<br />

in Mexico from those in <strong>the</strong> U.S. Concomitantly,<br />

land ownership patterns differ and influence<br />

natural resource management. <strong>Mexican</strong><br />

lands are classified into three types of tenancy:<br />

1. Federal lands include all lands administered<br />

under Federal Government institutions.<br />

Federal lands include protected<br />

natural areas such as Reserves of <strong>the</strong><br />

Biosphere, National Parks, and Areas of<br />

Protection of Natural Resources. Protected<br />

natural areas comprise 3% of <strong>the</strong><br />

total area of Mexico’s five recovery units.<br />

Volume I/Part I<br />

10<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

2. Ejidal lands are allotted by <strong>the</strong> <strong>Mexican</strong><br />

Government to a person or community<br />

<strong>for</strong> agriculture, <strong>for</strong>estry, mining, and<br />

o<strong>the</strong>r uses. Thus, lands within ejidos are<br />

intensively managed <strong>for</strong> natural resource<br />

use. Ejidos comprise approximately<br />

17.5% of <strong>the</strong> area within <strong>the</strong> <strong>Mexican</strong><br />

recovery units.<br />

3. Private lands are possessed under a<br />

“certificate of inaffectability.” Any<br />

protection af<strong>for</strong>ded <strong>the</strong>se lands is at <strong>the</strong><br />

discretion of <strong>the</strong> landowner. Approximately<br />

79.5% of <strong>the</strong> <strong>Mexican</strong> recovery<br />

units is comprised of private land.


This section describes various considerations,<br />

o<strong>the</strong>r than <strong>the</strong> basic biology of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl, that were integral in<br />

development of this <strong>Recovery</strong> <strong>Plan</strong>.<br />

Volume I/Part I<br />

D. D. CONSIDERATIONS CONSIDERATIONS IN<br />

IN<br />

RECOVERY RECOVERY PLAN PLAN DEVELOPMENT<br />

DEVELOPMENT<br />

RECOVERY RECOVERY UNITS<br />

UNITS<br />

The <strong>Mexican</strong> spotted owl is a widespread<br />

subspecies that occurs in a wide variety of<br />

habitats (see Part II). In addition, <strong>the</strong> threats<br />

faced by <strong>the</strong> subspecies, <strong>the</strong> management regimes<br />

employed by various agencies and in each<br />

country, and <strong>the</strong> protective mechanisms available<br />

in different portions of <strong>the</strong> subspecies’ range are<br />

variable. Finally, spotted owl densities, food<br />

habits, degree of isolation, and o<strong>the</strong>r aspects of<br />

<strong>the</strong> subspecies’ biology differ somewhat among<br />

portions of its range. For <strong>the</strong>se reasons, <strong>the</strong><br />

<strong>Recovery</strong> Team partitioned <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl range into distinct recovery units. Six<br />

recovery units were designated in <strong>the</strong> United<br />

States: Colorado Plateau, Sou<strong>the</strong>rn Rocky<br />

Mountains - Colorado, Sou<strong>the</strong>rn Rocky Mountains<br />

- New Mexico, Upper Gila Mountains,<br />

Basin and Range - West, and Basin and Range -<br />

East (Figs. II.B.1 and II.B.3–II.B.8). Five recovery<br />

units were established in Mexico: Sierra<br />

Madre Occidental - Norte, Sierra Madre Occidental<br />

- Sur, Sierra Madre Oriental - Norte,<br />

Sierra Madre Oriental - Sur, and Eje<br />

Neovolcanico (Figs. II.B.2). For a complete<br />

description of <strong>the</strong> recovery units and <strong>the</strong> bases<br />

<strong>for</strong> <strong>the</strong>ir designation see II.B.<br />

Whereas some management recommendations<br />

apply to <strong>the</strong> subspecies rangewide, delineating<br />

recovery units allowed specific recommendations<br />

to be prioritized appropriately within<br />

each portion of <strong>the</strong> subspecies’ range. In addition,<br />

some criteria <strong>for</strong> delisting <strong>the</strong> subspecies<br />

apply at <strong>the</strong> recovery-unit level. This approach<br />

allows delisting of <strong>the</strong> <strong>Mexican</strong> spotted owl by<br />

recovery unit when certain rangewide population<br />

and habitat criteria are met and when regional<br />

management plans or o<strong>the</strong>r sufficient regulatory<br />

mechanisms are implemented.<br />

11<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

THE THE CURRENT CURRENT SITUATION<br />

SITUATION<br />

In developing this <strong>Recovery</strong> <strong>Plan</strong>, <strong>the</strong> <strong>Recovery</strong><br />

Team considered various aspects of <strong>the</strong><br />

current spotted owl population, habitat, and<br />

threats. Two salient points emerged. First, <strong>the</strong><br />

<strong>Recovery</strong> Team assumes that <strong>the</strong> current population<br />

size and distribution are adequate <strong>for</strong><br />

providing a reference point <strong>for</strong> assessing future<br />

changes in <strong>the</strong> population, since no undisputable<br />

evidence is available indicating that <strong>the</strong> population<br />

is declining or is significantly below historical<br />

levels. This is a critical assumption that must<br />

be tested through <strong>the</strong> population monitoring<br />

required by this <strong>Recovery</strong> <strong>Plan</strong>. If <strong>the</strong> monitoring<br />

data demonstrate that <strong>the</strong> population is<br />

stable or increasing, <strong>the</strong> assumption of adequate<br />

population size will be validated. Conversely, if<br />

monitoring data show a decreasing population,<br />

<strong>the</strong> situation will need to be reexamined and<br />

corrective measures must be developed. Thus,<br />

<strong>the</strong> population and habitat monitoring requirements<br />

are essential parts of this <strong>Recovery</strong> <strong>Plan</strong>; if<br />

<strong>the</strong>se monitoring ef<strong>for</strong>ts are not conducted, <strong>the</strong><br />

management recommendations provided herein<br />

cannot stand alone.<br />

A second consideration involves variations in<br />

both spotted owl densities and threats faced<br />

throughout <strong>the</strong> subspecies’ range. <strong>Spotted</strong> owl<br />

densities are greatest in <strong>the</strong> center of <strong>the</strong> subspecies’<br />

range and <strong>the</strong>y decrease toward <strong>the</strong> range<br />

periphery. In addition, <strong>the</strong> main threats identified<br />

during <strong>the</strong> listing process were <strong>for</strong>estry<br />

practices and wildfire risk, both of which vary<br />

across <strong>the</strong> subspecies’ range. Table I.D.1. illustrates<br />

<strong>the</strong> <strong>Recovery</strong> Team’s appraisal.<br />

The Upper Gila Mountains, Basin and<br />

Range - West, and Basin and Range - East<br />

<strong>Recovery</strong> Units have significant owl populations<br />

with <strong>the</strong> potential of being seriously impacted by<br />

fire and/or <strong>for</strong>estry practices (Table I.D.1). This<br />

conclusion does not imply that <strong>the</strong> o<strong>the</strong>r recovery<br />

units are not important, but leads to <strong>the</strong><br />

recommendations that (1) recovery ef<strong>for</strong>ts<br />

concentrate on recovery units with <strong>the</strong> highest<br />

owl populations and where significant threats


exist; (2) management within recovery units<br />

should emphasize alleviating <strong>the</strong> greatest threats;<br />

and (3) <strong>the</strong> management recommendations in<br />

Part III should be tailored to <strong>the</strong> owl population<br />

and <strong>the</strong> threats existing in <strong>the</strong> specific area under<br />

analysis.<br />

The <strong>Recovery</strong> Team believes <strong>the</strong> risk of<br />

extirpation of <strong>Mexican</strong> spotted owls under <strong>the</strong><br />

near-term management recommendations is low.<br />

This belief is based on two points:<br />

1. Implementation of <strong>the</strong> management<br />

recommendations within this <strong>Recovery</strong><br />

<strong>Plan</strong> (see Part III) will protect occupied<br />

habitat, protect o<strong>the</strong>r habitat that can<br />

presumably be occupied in <strong>the</strong> near<br />

future, and allow <strong>for</strong> “replacement”<br />

habitat to develop and be sustained on<br />

<strong>the</strong> landscape. Habitat monitoring as<br />

required by this <strong>Recovery</strong> <strong>Plan</strong> should<br />

provide data on habitat trends throughout<br />

<strong>Recovery</strong> <strong>Plan</strong> duration.<br />

2. The population will be monitored over<br />

<strong>the</strong> life of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>, thus<br />

providing insight as to whe<strong>the</strong>r <strong>the</strong><br />

current “baseline” population is sufficient<br />

to maintain <strong>the</strong> subspecies over<br />

time and testing <strong>the</strong> assumption that<br />

<strong>the</strong> “baseline” population is adequate.<br />

The <strong>Recovery</strong> Team did not make <strong>the</strong><br />

assumption that <strong>the</strong> “baseline” population<br />

is adequate lightly, but reasoned<br />

that <strong>the</strong> <strong>Mexican</strong> spotted owl is well<br />

distributed throughout its historical<br />

range, suggesting that no significant<br />

extirpations have occurred.<br />

Volume I/Part I<br />

12<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able I.D.1. I.D.1. Summary of relative <strong>Mexican</strong> spotted owl population size, timber harvest threat, and<br />

fire threat by U.S. <strong>Recovery</strong> Unit.<br />

Thr Threat Thr eat S SSignificance<br />

S ignificance<br />

Reco eco eco ecover eco er ery er y U UUnit<br />

U Unit<br />

nit Population opulation Fir ir ir ire ir Timber imber<br />

Colorado Plateau low moderate low<br />

Sou<strong>the</strong>rn Rocky Mtns-CO low high low<br />

Sou<strong>the</strong>rn Rocky Mtns-NM low high high<br />

Upper Gila Mountains high high high<br />

Basin and Range - West high high low<br />

Basin and Range - East high high high<br />

RECOVERY RECOVERY PLAN PLAN DURATION<br />

DURATION<br />

Any management plan must specify <strong>the</strong> time<br />

period over which <strong>the</strong> plan is to be implemented.<br />

The <strong>Recovery</strong> Team decided that a 10year<br />

period is appropriate <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

<strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong> (assuming <strong>the</strong><br />

delisting criteria specified in III.A are met) <strong>for</strong><br />

several reasons:<br />

1. Ten years allows adequate time to monitor<br />

<strong>the</strong> trends in population and habitat.<br />

The charge of <strong>the</strong> <strong>Recovery</strong> Team was<br />

to develop a plan that would lead to<br />

recovery of <strong>the</strong> subspecies. In developing<br />

<strong>the</strong> delisting criteria specified in III.A,<br />

<strong>the</strong> <strong>Recovery</strong> Team reasoned that <strong>the</strong><br />

population must be stable or increasing<br />

be<strong>for</strong>e <strong>the</strong> subspecies could be considered<br />

<strong>for</strong> delisting. The <strong>Recovery</strong> Team<br />

fur<strong>the</strong>r determined that a monitoring<br />

period of 10 years would provide in<strong>for</strong>mation<br />

about population trends that<br />

could be used with a reasonably high<br />

level of confidence. The five-year monitoring<br />

period <strong>the</strong> Act requires after a<br />

species is delisted will fur<strong>the</strong>r increase<br />

confidence in trend in<strong>for</strong>mation.<br />

2. A 10-year period should be sufficient<br />

time to fill some of <strong>the</strong> major gaps in<br />

existing knowledge, and accommodate<br />

possible changes in future conditions.<br />

Many aspects of <strong>Mexican</strong> spotted owl<br />

biology remain unknown or poorly<br />

understood. Consequently, <strong>the</strong> effects of


Volume I/Part I<br />

different resource-management practices<br />

on <strong>the</strong> fitness of individuals and on<br />

population persistence remain unclear. A<br />

better understanding of <strong>the</strong>se relationships<br />

is needed be<strong>for</strong>e a viable long-term<br />

management plan can be developed.<br />

Implementing <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> includes<br />

conducting <strong>the</strong> research activities<br />

recommended in III.D; if <strong>the</strong>se studies<br />

are started immediately, 10 years should<br />

be adequate to complete <strong>the</strong> majority of<br />

<strong>the</strong>m.<br />

3. Uncertainty about <strong>the</strong> future could<br />

render this <strong>Recovery</strong> <strong>Plan</strong> inadequate,<br />

unacceptable, or o<strong>the</strong>rwise obsolete.<br />

Future events and developments could<br />

have social, economic, environmental,<br />

and o<strong>the</strong>r ramifications that cannot be<br />

predicted. To try to plan beyond <strong>the</strong> next<br />

decade or so would require an unjustified<br />

confidence in our ability to predict <strong>the</strong><br />

state of our society and <strong>the</strong> environment.<br />

4. Consistency with <strong>the</strong> requirements of <strong>the</strong><br />

Act. The Act requires that <strong>the</strong> status of<br />

listed species be reviewed every five years.<br />

This <strong>Recovery</strong> <strong>Plan</strong> constitutes an in–<br />

depth status review of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, so a <strong>for</strong>mal status review<br />

should be conducted in years five and 10<br />

of <strong>Recovery</strong> <strong>Plan</strong> implementation.<br />

Unless new in<strong>for</strong>mation or o<strong>the</strong>r developments<br />

render this <strong>Recovery</strong> <strong>Plan</strong><br />

obsolete in <strong>the</strong> interim, <strong>the</strong> 10-year<br />

point should mark <strong>the</strong> end of this<br />

<strong>Recovery</strong> <strong>Plan</strong> and implementation of a<br />

longer-term management strategy.<br />

Several reviewers of <strong>the</strong> draft version of this<br />

<strong>Recovery</strong> <strong>Plan</strong> pointed out that this relatively<br />

short <strong>Recovery</strong> <strong>Plan</strong> duration fails to take into<br />

account <strong>the</strong> long-term processes that have<br />

influenced and will continue to influence<br />

dynamic ecosystems. The <strong>Recovery</strong> Team believes,<br />

however, that <strong>the</strong> management recommended<br />

<strong>for</strong> <strong>the</strong> next few years was developed<br />

with consideration of <strong>the</strong> long- and short-term<br />

effects of <strong>the</strong>se near-term management recommendations.<br />

13<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Based on <strong>the</strong> <strong>for</strong>egoing points, <strong>the</strong> <strong>Recovery</strong><br />

Team recommends a <strong>Recovery</strong> <strong>Plan</strong> duration of<br />

10 years unless data indicate that earlier revision<br />

is appropriate, or that <strong>the</strong> applicable recommendations<br />

be continued beyond that time. The<br />

monitoring and research to be conducted during<br />

<strong>the</strong> life of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> will resolve much<br />

uncertainty surrounding <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl. The uncertainty about <strong>the</strong> future can never<br />

be resolved, but a better understanding of<br />

<strong>Mexican</strong> spotted owl natural history will enhance<br />

our ability to create a long-term management<br />

plan.<br />

CONSERVATION CONSERVATION PLANS<br />

PLANS<br />

FOR FOR OTHER OTHER OTHER SPOTTED SPOTTED OWL<br />

OWL<br />

SUBSPECIES<br />

SUBSPECIES<br />

SUBSPECIES<br />

Several conservation strategies have been<br />

developed <strong>for</strong> <strong>the</strong> o<strong>the</strong>r spotted owl subspecies.<br />

Perhaps <strong>the</strong> best known subspecies is <strong>the</strong> nor<strong>the</strong>rn<br />

spotted owl of <strong>the</strong> Pacific Northwest and<br />

northwestern Cali<strong>for</strong>nia. The nor<strong>the</strong>rn subspecies<br />

was listed as threatened in June 1990,<br />

resulting in extensive conflict between conservation<br />

of <strong>the</strong> subspecies and economic and social<br />

interests of <strong>the</strong> Pacific Northwest, particularly<br />

<strong>the</strong> timber industry.<br />

The first management strategy was initiated<br />

by <strong>the</strong> FS in <strong>the</strong> late 1970s. That approach,<br />

which continued within some portions of <strong>the</strong><br />

subspecies’ range until 1990, was to manage<br />

individual spotted owl territories, called <strong>Spotted</strong><br />

<strong>Owl</strong> Habitat Areas (SOHAs), or, earlier, <strong>Spotted</strong><br />

<strong>Owl</strong> Management Areas. Each SOHA consisted<br />

of certain acreages that varied according to<br />

location. Those territories were established<br />

according to certain clustering and spacing<br />

guidelines, and <strong>the</strong> general prescription <strong>for</strong> <strong>the</strong><br />

territories was to restrict timber harvest so that a<br />

minimum “suitable habitat” acreage standard was<br />

maintained in <strong>the</strong> territories. However, in certain<br />

circumstances some harvest was allowed, such as<br />

salvage harvest.<br />

In 1989, in response to increasing controversy<br />

over <strong>the</strong> spotted owl issue, <strong>the</strong> difficulty<br />

<strong>the</strong> issue was causing land-management agencies,<br />

and <strong>the</strong> proposed listing of <strong>the</strong> nor<strong>the</strong>rn spotted<br />

owl as a threatened species, <strong>the</strong> Interagency<br />

Scientific Committee (ISC) was established. The


ISC produced “A Conservation Strategy <strong>for</strong> <strong>the</strong><br />

Nor<strong>the</strong>rn <strong>Spotted</strong> <strong>Owl</strong>” (Thomas et al. 1990),<br />

which recommended significant changes in<br />

spotted owl management on public lands in <strong>the</strong><br />

Pacific Northwest. Briefly, <strong>the</strong> ISC delineated<br />

large blocks of owl habitat, called Habitat<br />

Conservation Areas (HCAs). The goal was to<br />

delineate, where possible, HCAs known to<br />

support or with <strong>the</strong> potential to support at least<br />

20 spotted owl pairs. However, where 20-pair<br />

HCAs were not possible, HCAs of 1 to 19 pairs<br />

were delineated. The HCAs were spaced certain<br />

distances from one ano<strong>the</strong>r, depending on <strong>the</strong>ir<br />

sizes. The HCAs were to be managed so that no<br />

habitat degradation occurred within <strong>the</strong>m,<br />

protecting existing habitat and allowing previously<br />

disturbed areas to return to a suitable<br />

condition. The ISC envisioned eventual HCAs<br />

where owl pairs could freely interact without<br />

significant disruption of habitat continuity<br />

between territories.<br />

In addition, <strong>the</strong> ISC recommended managing<br />

<strong>the</strong> areas between HCAs, termed <strong>the</strong> “<strong>for</strong>est<br />

matrix,” according to <strong>the</strong> “50-11-40 rule.” This<br />

rule prescribed that at least 50% of <strong>the</strong> <strong>for</strong>ested<br />

area within each quarter-township was to contain<br />

trees averaging a minimum of 28 cm (11 in)<br />

in diameter and with at least 40% crown closure.<br />

The idea was that <strong>the</strong>se conditions would allow<br />

movement of owls between HCAs, <strong>the</strong>reby<br />

allowing genetic flow and demographic rescue of<br />

subpopulations. The ISC also recommended<br />

retention of 28-ha (70-acre) areas within <strong>the</strong><br />

<strong>for</strong>est matrix to possibly provide future nesting/<br />

roosting sites.<br />

In 1991, <strong>the</strong> Secretary of <strong>the</strong> Interior appointed<br />

<strong>the</strong> Nor<strong>the</strong>rn <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

Team and charged it with developing a recovery<br />

plan <strong>for</strong> that subspecies. That recovery plan was<br />

closely modeled after <strong>the</strong> ISC plan. The HCA<br />

network was modified based on updated in<strong>for</strong>mation,<br />

resulting in a network of Designated<br />

Conservation Areas (DCAs). Timber harvest in<br />

DCAs was generally not allowed in suitable<br />

habitat. Silvicultural treatments designed to<br />

encourage spotted owl habitat were limited to no<br />

more than 5% of a DCA in <strong>the</strong> first five years of<br />

plan implementation. Management recommendations<br />

<strong>for</strong> areas outside DCAs deviated from<br />

<strong>the</strong> ISC approach by providing greater justifica-<br />

Volume I/Part I<br />

14<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

tion <strong>for</strong> specific recommendations and consideration<br />

of economic efficiency of implementation.<br />

The Nor<strong>the</strong>rn <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

was never implemented. Instead, <strong>the</strong> most recent<br />

management strategy <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted<br />

owl resulted from analyses and recommendations<br />

<strong>for</strong>mulated by <strong>the</strong> Forest Ecosystem Management<br />

Assessment Team (FEMAT). The FEMAT<br />

was appointed by President Clinton to develop<br />

alternative plans <strong>for</strong> management of late-successional<br />

ecosystem components including, but not<br />

specific to, <strong>the</strong> nor<strong>the</strong>rn spotted owl. Only<br />

Federal land management was addressed in <strong>the</strong><br />

FEMAT alternatives.<br />

The President selected “Option 9” developed<br />

by <strong>the</strong> FEMAT, which calls <strong>for</strong> a series of Late<br />

Successional Reserves (LSRs) corresponding<br />

roughly to <strong>the</strong> HCAs under <strong>the</strong> ISC plan.<br />

Outside <strong>the</strong> LSRs, <strong>the</strong> <strong>for</strong>est matrix includes<br />

Riparian Reserves (various sized buffers along<br />

class 1-3 streams); green-tree retention requirements<br />

(where 15% of each watershed is managed<br />

<strong>for</strong> late successional <strong>for</strong>est and at least 15% of<br />

each harvest unit is retained in <strong>the</strong> latest successional<br />

<strong>for</strong>est available); and preservation of 40 ha<br />

(100 ac) around all owl sites known as of 1<br />

January 1994. The goal of this matrix prescription<br />

is to accommodate dispersing and “floater”<br />

owls, as well as o<strong>the</strong>r species dependent on old<br />

and mature <strong>for</strong>est conditions.<br />

The FEMAT plan also establishes Adaptive<br />

Management Areas (AMAs) in Cali<strong>for</strong>nia,<br />

Oregon, and Washington. These AMAs vary<br />

from less than 40,500 ha (100,000 ac) to nearly<br />

200,000 ha (500,000 ac). The management<br />

objectives <strong>for</strong> each AMA also vary, but <strong>the</strong>y are<br />

generally established to develop and test techniques<br />

<strong>for</strong> active <strong>for</strong>est management that provide<br />

a wide range of resource values including <strong>for</strong>est<br />

products, late-successional <strong>for</strong>est habitat, and<br />

high-quality recreation.<br />

The range of <strong>the</strong> Cali<strong>for</strong>nia spotted owl<br />

abuts <strong>the</strong> range of <strong>the</strong> nor<strong>the</strong>rn subspecies in<br />

nor<strong>the</strong>astern Cali<strong>for</strong>nia, ranging south through<br />

<strong>the</strong> Sierra Nevada, west through <strong>the</strong> “transverse<br />

ranges” of Sou<strong>the</strong>rn Cali<strong>for</strong>nia, <strong>the</strong>n north along<br />

<strong>the</strong> Coast Range to Monterey County. The<br />

Cali<strong>for</strong>nia spotted owl was first managed by <strong>the</strong><br />

FS using <strong>the</strong> SOHA system described <strong>for</strong> <strong>the</strong><br />

nor<strong>the</strong>rn subspecies. The portion of <strong>the</strong> subspe-


cies’ range in Sou<strong>the</strong>rn Cali<strong>for</strong>nia is still managed<br />

under that system while its effectiveness is<br />

assessed. However, <strong>the</strong> current management<br />

regime in <strong>the</strong> Sierra Nevada is described in <strong>the</strong><br />

“Assessment of <strong>the</strong> Current Status of <strong>the</strong> Cali<strong>for</strong>nia<br />

<strong>Spotted</strong> <strong>Owl</strong>, with Recommendations <strong>for</strong><br />

Management” (Caspow <strong>Plan</strong>) (Verner et al.<br />

1992b). The FS implemented <strong>the</strong> Caspow <strong>Plan</strong><br />

on a temporary basis through an environmental<br />

assessment under <strong>the</strong> National Environmental<br />

Policy Act (NEPA) on 1 March 1993. The FS<br />

has since released a draft environmental impact<br />

statement under <strong>the</strong> NEPA that analyzes several<br />

alternatives <strong>for</strong> Cali<strong>for</strong>nia spotted owl management,<br />

one of which is <strong>the</strong> Caspow <strong>Plan</strong>.<br />

In <strong>the</strong> Sierra Nevada, <strong>the</strong> Caspow <strong>Plan</strong><br />

recommends management emphasizing adequate<br />

amounts and distribution of suitable owl habitat<br />

through improved habitat and resource management<br />

practices ra<strong>the</strong>r than through protection of<br />

large blocks of habitat. The strategy seeks to<br />

protect known owl nest or roost sites, retain <strong>the</strong><br />

larger and older components of <strong>for</strong>est structure,<br />

and address <strong>the</strong> problems of fire suppression and<br />

fuel loading. Additional objectives include<br />

rapidly recovering nesting and roosting habitat<br />

following disturbance, maintaining existing<br />

canopy layers, promoting tree growth by thinning<br />

in middle and lower canopy layers, and<br />

reducing vertical fuel ladders. Habitat available<br />

<strong>for</strong> timber management is classified by structural<br />

condition and utility <strong>for</strong> various life history<br />

requirements, and <strong>the</strong> resultant habitat classes<br />

are managed with restrictions on structural<br />

modifications. Long-term management proposals<br />

also focus on spotted owl nesting and roosting<br />

habitat as <strong>the</strong> target conditions <strong>for</strong> silvicultural<br />

activities.<br />

To <strong>for</strong>mulate <strong>the</strong> management strategy<br />

contained in this <strong>Recovery</strong> <strong>Plan</strong>, <strong>the</strong> <strong>Recovery</strong><br />

Team examined <strong>the</strong> extensive ef<strong>for</strong>ts to protect<br />

<strong>the</strong> Cali<strong>for</strong>nia and nor<strong>the</strong>rn spotted owl subspecies.<br />

These ef<strong>for</strong>ts have, so far, experienced<br />

varying degrees of success and controversy.<br />

Moreover, <strong>the</strong> three subspecies exhibit differences<br />

in habitat use, habitat distribution, and<br />

threats. The <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

<strong>Plan</strong> combines, in <strong>the</strong> <strong>Recovery</strong> Team’s opinion,<br />

<strong>the</strong> applicable recommendations from o<strong>the</strong>r<br />

Volume I/Part I<br />

15<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

planning ef<strong>for</strong>ts with those uniquely applicable<br />

to <strong>the</strong> <strong>Mexican</strong> subspecies.<br />

ECOSYSTEM ECOSYSTEM MANAGEMENT<br />

MANAGEMENT<br />

The development of ecosystem management<br />

in <strong>the</strong> history of conservation plans <strong>for</strong> <strong>the</strong><br />

nor<strong>the</strong>rn spotted owl was described by Meslow<br />

(1993). Results of population and habitat<br />

research were incorporated into landscape<br />

designs involving reserves (HCAs) and interstitial<br />

matrices, both of which are basic attributes<br />

of landscape ecology (Diaz and Apostol 1993).<br />

Fur<strong>the</strong>r considerations of <strong>the</strong> ecosystem management<br />

approach were extended to sympatric<br />

Federal- and State-listed species. It was clear that<br />

single-species management <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn<br />

spotted owl would have numerous impacts on<br />

<strong>the</strong> many eligible but yet to be listed species<br />

(USDI 1992; Block et al. 1995). Thomas et al.<br />

(1993) described <strong>the</strong> relationship between <strong>the</strong><br />

ISC plan <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted owl and <strong>the</strong><br />

likelihood of viability <strong>for</strong> a suite of o<strong>the</strong>r species<br />

closely associated with late-successional <strong>for</strong>est.<br />

Verner et al. (1992) described numerous links<br />

between <strong>the</strong> Cali<strong>for</strong>nia spotted owl and associated<br />

ecosystem components that it uses and<br />

requires to survive. Assessments of o<strong>the</strong>r species,<br />

such as nor<strong>the</strong>rn goshawks (Reynolds et al.<br />

1992) have also underscored <strong>the</strong> need to manage<br />

large landscapes to provide adequate prey and<br />

<strong>the</strong> diversity of habitats needed by those species.<br />

Despite growing academic and professional<br />

awareness of <strong>the</strong> need to manage entire ecosystems,<br />

<strong>the</strong> <strong>Recovery</strong> Team is charged with development<br />

of a recovery plan <strong>for</strong> a single species.<br />

However, as <strong>the</strong> FWS and o<strong>the</strong>r land-management<br />

agencies move toward managing entire<br />

ecosystems, <strong>the</strong>y are recognizing that singlespecies<br />

management will never protect all of <strong>the</strong><br />

organisms that comprise <strong>the</strong> ecosystems upon<br />

which target species depend. Fur<strong>the</strong>rmore, a<br />

management plan <strong>for</strong> one species may conflict<br />

with a management plan <strong>for</strong> a sympatric species<br />

in absence of careful integration of <strong>the</strong> two<br />

plans. Block and Brennan (1993) noted that <strong>the</strong><br />

management recommendations of <strong>the</strong> ISC <strong>for</strong><br />

<strong>the</strong> nor<strong>the</strong>rn spotted owl were firmly based in<br />

habitat management. In addition to habitat,<br />

however, both ecosystem-oriented and popula-


tion-level considerations must be wed in conservation<br />

planning (Gutiérrez 1994).<br />

The recovery team considered <strong>the</strong> interaction<br />

of populations, habitats, and ecosystems in<br />

<strong>the</strong> development of this <strong>Recovery</strong> <strong>Plan</strong>. The<br />

<strong>Recovery</strong> Team recognizes that numerous habitats<br />

exist within <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl and that not all of those habitats are<br />

important to <strong>the</strong> subspecies. The <strong>Recovery</strong> Team<br />

concentrated its management recommendations<br />

on habitats known to be important to <strong>the</strong> owl<br />

(see Parts II and III), while allowing o<strong>the</strong>r<br />

ecosystem management objectives, such as<br />

conservation of o<strong>the</strong>r species, to drive management<br />

of habitats where spotted owls are a secondary<br />

concern.<br />

The <strong>Recovery</strong> Team believes that it is important<br />

to evaluate <strong>the</strong> effects of implementing<br />

<strong>Recovery</strong> <strong>Plan</strong> management recommendations<br />

on o<strong>the</strong>r endangered, threatened, sensitive,<br />

candidate, or o<strong>the</strong>r species of concern. In addition,<br />

it is important that <strong>the</strong> recommendations<br />

<strong>for</strong> <strong>Mexican</strong> spotted owl management be compared<br />

with <strong>the</strong> recommendations in o<strong>the</strong>r<br />

species’ recovery or management plans. If conflicts<br />

are identified, <strong>the</strong>y need to be resolved by<br />

appropriate land managers and/or scientists.<br />

An important objective in management of<br />

<strong>for</strong>ested ecosystems should be to address <strong>for</strong>est<br />

health problems, return <strong>for</strong>ested ecosystems to<br />

conditions within <strong>the</strong>ir natural range of variation,<br />

and work toward sustainable and resilient<br />

ecosystems. The goals of this <strong>Recovery</strong> <strong>Plan</strong> and<br />

ecosystem management principles are compatible.<br />

Proper ecosystem management will provide<br />

<strong>for</strong> landscapes in which spotted owls and o<strong>the</strong>r<br />

ecosystem components persist within <strong>the</strong> range<br />

of <strong>the</strong>ir evolutionary adaptations. The metric<br />

used to measure progress should be <strong>the</strong> amount<br />

of acreage successfully treated to meet a desired<br />

result, and not commodity-based measures such<br />

as “board feet” or “animal unit months.” Commodities<br />

will undoubtedly be a byproduct of<br />

<strong>for</strong>ested ecosystem management, but should not<br />

be <strong>the</strong> driving consideration.<br />

ECONOMIC ECONOMIC CONSIDERATIONS<br />

CONSIDERATIONS<br />

Protecting threatened and endangered<br />

species can conflict with o<strong>the</strong>r resource objec-<br />

Volume I/Part I<br />

16<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

tives. These conflicts can become more intense<br />

when species conservation ef<strong>for</strong>ts restrict economic<br />

returns from lands people depend upon<br />

<strong>for</strong> <strong>the</strong>ir livelihoods and communities depend<br />

upon <strong>for</strong> <strong>the</strong>ir very existence. Whe<strong>the</strong>r conflicts<br />

between species conservation and economic<br />

return are real or perceived, human concerns<br />

should be considered so long as <strong>the</strong> conservation<br />

goal is achieved.<br />

As mentioned previously, <strong>the</strong> <strong>Recovery</strong><br />

Team’s charge was to develop a plan that would<br />

lead to recovery of <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

However, specific cause-effect relationships of<br />

many management activities on individual owls<br />

and pairs or in relation to population processes<br />

are not entirely clear. Given <strong>the</strong>se uncertainties,<br />

it may be tempting to take a conservative approach<br />

to recovery by recommending cessation<br />

of all anthropogenic activities <strong>for</strong> which effects<br />

of <strong>the</strong> activity on <strong>the</strong> target species are poorly<br />

understood. However, recommendations <strong>for</strong><br />

resource management should be based on<br />

established in<strong>for</strong>mation. The absence of needed<br />

in<strong>for</strong>mation should stimulate research, and <strong>the</strong><br />

results of that research should guide management.<br />

The only way to understand <strong>the</strong> causeand-effect<br />

relationships between management<br />

actions and specific resources is by studying<br />

<strong>the</strong>m, preferably through controlled experiments.<br />

The recommendations contained herein<br />

allow most land-management activities to occur<br />

provided that <strong>the</strong> effects of those activities are<br />

evaluated during <strong>the</strong> recovery period. In addition,<br />

<strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> recommends that<br />

scientific monitoring of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

population and its habitat should accompany<br />

those activities to assess <strong>the</strong>ir impact on spotted<br />

owl populations. If warranted, <strong>the</strong>se activities<br />

can be altered or eliminated if monitoring or<br />

research indicates a significant risk to <strong>the</strong> spotted<br />

owl population. In o<strong>the</strong>r cases, restrictions on<br />

human activities are recommended where data<br />

show a high likelihood that <strong>the</strong> spotted owl’s<br />

persistence may be significantly compromised if<br />

certain land-management practices continue.<br />

Obviously, <strong>the</strong> decision on which activities<br />

must be altered or eliminated and which may<br />

proceed if closely monitored cannot be made<br />

with absolute certainty. Such decisions require


professional judgement of <strong>the</strong> most qualified<br />

scientists using <strong>the</strong> best available data. The FWS<br />

selected <strong>the</strong> <strong>Recovery</strong> Team members with that<br />

fact in mind. However, <strong>the</strong> FWS also intends to<br />

use <strong>the</strong> expertise of o<strong>the</strong>rs who may contribute<br />

useful in<strong>for</strong>mation to improve management of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Any conservation plan, regardless of species,<br />

must include <strong>the</strong> considerations discussed above<br />

when uncertainties exist. The FWS is confident<br />

that this <strong>Recovery</strong> <strong>Plan</strong>, given its inherent<br />

flexibility, has a high likelihood of leading to <strong>the</strong><br />

recovery of <strong>the</strong> <strong>Mexican</strong> spotted owl without<br />

causing unacceptable levels of economic and<br />

social hardship during its implementation.<br />

HUMAN HUMAN INTERVENTION<br />

INTERVENTION<br />

AND AND NATURAL NATURAL NATURAL PROCESSES<br />

PROCESSES<br />

Much criticism directed at <strong>Mexican</strong> spotted<br />

owl management centers on <strong>the</strong> concept that<br />

today’s southwestern <strong>for</strong>ests are in an unnatural<br />

state; that grazing, fire suppression, <strong>for</strong>estry<br />

practices, and o<strong>the</strong>r anthropogenic processes<br />

have led to <strong>for</strong>est conditions much denser than<br />

those existing during presettlement times. A<br />

concurrent increase in mixed-conifer <strong>for</strong>ests is<br />

also believed to have occurred. These points<br />

lead some to <strong>the</strong> conclusion that <strong>the</strong> <strong>Mexican</strong><br />

spotted owl population is at an all-time (and<br />

unsustainable) high. The <strong>Recovery</strong> Team is<br />

unaware of data that clearly support that conclusion,<br />

and questions whe<strong>the</strong>r stands recently<br />

converted to mixed-conifer <strong>for</strong>est possess <strong>the</strong><br />

structural characteristics utilized by <strong>the</strong> subspecies.<br />

The <strong>Recovery</strong> Team acknowledges that<br />

humans have had a pronounced influence on<br />

contemporary <strong>for</strong>est conditions; however, <strong>the</strong><br />

effects of human activities on <strong>the</strong> <strong>Mexican</strong><br />

spotted owl population are unknown. Even if<br />

one accepts that <strong>the</strong> spotted owl population in<br />

mixed-conifer <strong>for</strong>est is unnaturally high, one<br />

must also consider that o<strong>the</strong>r habitats that may<br />

have been important historically, such as lowerand<br />

middle-elevation riparian areas, have been<br />

dramatically reduced. These two trends may be<br />

offsetting, and <strong>the</strong> net gain or loss of spotted owl<br />

carrying capacity can only be speculated upon.<br />

Volume I/Part I<br />

17<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

It would be imprudent, if not impossible, to<br />

develop a management plan <strong>for</strong> a species by<br />

speculating on its status in <strong>the</strong> distant past;<br />

ra<strong>the</strong>r, <strong>the</strong> appropriate approach is to acknowledge<br />

that we are dealing with a drastically altered<br />

landscape and that a return to presettlement<br />

conditions is impossible. In that light, <strong>the</strong><br />

<strong>Recovery</strong> Team acknowledges that humans have<br />

a major role to play in management of <strong>the</strong><br />

spotted owl and <strong>the</strong> <strong>for</strong>ests of <strong>the</strong> Southwest.<br />

The <strong>Recovery</strong> Team believes that a viable <strong>for</strong>estproducts<br />

industry is critical in carrying out <strong>the</strong><br />

management actions recommended in this<br />

<strong>Recovery</strong> <strong>Plan</strong>, making it an essential agent of<br />

plan implementation.<br />

DIFFERENCES DIFFERENCES BETWEEN<br />

BETWEEN<br />

THE THE DRAFT DRAFT AND AND FINAL FINAL<br />

FINAL<br />

RECOVERY RECOVERY RECOVERY PLANS<br />

PLANS<br />

The <strong>Recovery</strong> Team considered all comments<br />

received on <strong>the</strong> draft <strong>Mexican</strong> <strong>Spotted</strong><br />

<strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>. In addition, <strong>the</strong> draft<br />

<strong>Recovery</strong> <strong>Plan</strong> underwent extensive peer review<br />

from both purposely selected reviewers and<br />

“blind” reviewers selected by certain scientific<br />

societies (see Appendix E). These reviews led to a<br />

final <strong>Recovery</strong> <strong>Plan</strong> that differs substantially<br />

from <strong>the</strong> draft version. We do not attempt to<br />

detail every difference between <strong>the</strong> two versions<br />

of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>, but discuss <strong>the</strong>se differences<br />

in general terms.<br />

Part II of <strong>the</strong> draft <strong>Recovery</strong> <strong>Plan</strong> contained<br />

a great deal of technical in<strong>for</strong>mation. In <strong>the</strong><br />

interest of making <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> an easier<br />

document to use, several of those chapters were<br />

placed in a companion volume to this <strong>Recovery</strong><br />

<strong>Plan</strong>. The in<strong>for</strong>mation contained in those<br />

chapters was integral to <strong>Recovery</strong> <strong>Plan</strong> development,<br />

so <strong>the</strong> main points in each are summarized<br />

in Part II of this final <strong>Recovery</strong> <strong>Plan</strong>.<br />

Part III has changed substantially from <strong>the</strong><br />

draft version. Much of <strong>the</strong> background and<br />

justification discussion has been moved to Part<br />

II, so that <strong>the</strong> current Part III deals strictly with<br />

<strong>the</strong> management recommendations and delisting<br />

criteria. This allows land managers to more easily<br />

pull <strong>the</strong> specific recommendations out of <strong>the</strong><br />

Part III text. In addition, <strong>the</strong> management


ecommendations have changed considerably, in<br />

that <strong>the</strong>y are now in a more descriptive, ra<strong>the</strong>r<br />

than prescriptive, context. The changes were in<br />

response to comments from numerous land<br />

managers who expressed <strong>the</strong>ir concern that many<br />

of <strong>the</strong> tools available to achieve land-management<br />

objectives were overly constrained. The<br />

<strong>Recovery</strong> Team recognizes that <strong>the</strong> best approach<br />

is to describe <strong>the</strong> desired conditions on <strong>the</strong><br />

landscape, while providing land-management<br />

professionals <strong>the</strong> flexibility to choose <strong>the</strong> tools to<br />

achieve <strong>the</strong> stated objectives.<br />

Volume I/Part I<br />

18<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The most significant change in Part IV is<br />

that <strong>the</strong> responsibility <strong>for</strong> implementing some of<br />

<strong>the</strong> tasks recommended in this <strong>Recovery</strong> <strong>Plan</strong><br />

has been distributed among different entities. In<br />

addition, <strong>the</strong> estimated costs of implementing<br />

specific recovery tasks is provided.


Biological and Ecological Background<br />

Volume I, Part II


Volume I/Part II<br />

A. A. GENERAL GENERAL BIOLOGY<br />

BIOLOGY<br />

AND AND ECOLOGICAL ECOLOGICAL RELATIONSHIPS RELATIONSHIPS OF<br />

OF<br />

THE THE MEXICAN MEXICAN MEXICAN SPOTTED SPOTTED OWL<br />

OWL<br />

Sarah E. Rinkevich, Joseph L. Ganey, James P. Ward Jr.,<br />

Gary C. White, Dean L. Urban, Alan B. Franklin,<br />

William M. Block, and Fernando Clemente<br />

This section presents a summary of<br />

Volume 2, which examines aspects of <strong>the</strong> biology<br />

and ecological relationships of <strong>Mexican</strong> spotted<br />

owls in more detail. This is not an exhaustive<br />

treatment of <strong>the</strong> owl’s biology and ecology, but is<br />

intended to provide an overview of biological<br />

elements germane to recovering <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Although gaps still exist, our<br />

understanding of <strong>the</strong> <strong>Mexican</strong> spotted owl’s<br />

natural history has increased with recent research<br />

as well as data analyses accomplished by <strong>the</strong><br />

<strong>Recovery</strong> Team.<br />

A wealth of in<strong>for</strong>mation exists <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn<br />

and Cali<strong>for</strong>nia spotted owls (Thomas et al.<br />

1990, Bart et al. 1992, Verner et al. 1992a,<br />

Gutiérrez et al., 1995). Although different in<br />

some respects, many aspects of <strong>the</strong> owls’ biology<br />

and ecology are similar among <strong>the</strong> three subspecies.<br />

Thus, where appropriate, in<strong>for</strong>mation from<br />

<strong>the</strong>se subspecies was used <strong>for</strong> comparison or<br />

where data were limited regarding <strong>the</strong> <strong>Mexican</strong><br />

spotted owl.<br />

TAXONOMY<br />

TAXONOMY<br />

Three species within <strong>the</strong> genus <strong>Strix</strong> occur<br />

north of Mexico: spotted (S. <strong>occidentalis</strong>), barred<br />

(S. varia), and great gray owls (S. nebulosa).<br />

<strong>Mexican</strong> spotted, barred, and fulvous owls (S.<br />

fulvescens) occur in Mexico. The <strong>Mexican</strong> spotted<br />

owl (S. o. <strong>lucida</strong>) is one of three subspecies of<br />

spotted owl recognized by <strong>the</strong> American Ornithologists’<br />

Union (AOU) in its last checklist that<br />

included subspecies (AOU 1957:285). The o<strong>the</strong>r<br />

two subspecies are <strong>the</strong> nor<strong>the</strong>rn (S. o. caurina)<br />

and <strong>the</strong> Cali<strong>for</strong>nia spotted owl (S. o. <strong>occidentalis</strong>)<br />

(AOU 1957; Figure.II.A.1).<br />

The <strong>Mexican</strong> subspecies was first described<br />

from a specimen collected at Mount Tancitaro,<br />

Michoacan, Mexico and named Syrnium<br />

occidentale lucidum (Nelson 1903). The spotted<br />

owl was later assigned to <strong>the</strong> genus <strong>Strix</strong><br />

19<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

(Ridgway 1914) and <strong>the</strong> subspecific name was<br />

changed to <strong>lucida</strong> to con<strong>for</strong>m to taxonomic<br />

standards. Monson and Phillips (1981) regarded<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl in Arizona as S. o.<br />

huachucae, noting that <strong>the</strong>y were paler than S. o.<br />

<strong>lucida</strong>. However, this taxonomic designation was<br />

not followed by <strong>the</strong> AOU (1957).<br />

The <strong>Mexican</strong> subspecies is geographically<br />

isolated from both <strong>the</strong> Cali<strong>for</strong>nia and nor<strong>the</strong>rn<br />

subspecies. Using electrophoresis to examine<br />

allozyme variation, Barrowclough and Gutiérrez<br />

(1990) found a major allelic difference between<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl and <strong>the</strong> two coastal<br />

subspecies. This difference suggests that <strong>the</strong><br />

<strong>Mexican</strong> spotted owl has been isolated genetically<br />

from <strong>the</strong> o<strong>the</strong>r subspecies <strong>for</strong> considerable<br />

time, has followed a separate evolutionary<br />

history, and could <strong>the</strong>re<strong>for</strong>e be considered a<br />

separate species (Barrowclough and Gutiérrez<br />

1990:742).<br />

Nor<strong>the</strong>rn spotted owls are known to hybridize<br />

with barred owls. Hybrids have been found<br />

in Washington and Oregon (Hamer et al. 1992),<br />

and in Cali<strong>for</strong>nia (Alan Franklin, Humboldt<br />

State Univ., Arcata, CA, pers. comm.). The<br />

hybrids can be identified by <strong>the</strong>ir plumage,<br />

vocalizations, and morphology (Hamer et al.<br />

1992). Closely related species occasionally<br />

hybridize naturally, especially where habitat<br />

disruption has led to contact between species<br />

previously isolated geographically (Short 1965,<br />

1972). Hybridization has not been reported in<br />

<strong>the</strong> <strong>Mexican</strong> subspecies. The possibility of<br />

hybridization exists in Mexico where barred<br />

owls, fulvous owls, and spotted owls overlap in<br />

distribution. No evidence currently exists documenting<br />

actual sympatry among <strong>the</strong>se species,<br />

however.


Figur igur igure igur e II.A.1. II.A.1. Geographic range of <strong>the</strong> spotted owl.<br />

Volume I/Part II<br />

20<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Volume I/Part II<br />

DESCRIPTION<br />

DESCRIPTION<br />

The spotted owl is mottled in appearance<br />

with irregular white and brown spots on its<br />

abdomen, back, and head. The spots of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl are larger and more numerous<br />

than in <strong>the</strong> o<strong>the</strong>r two subspecies, giving it a<br />

lighter appearance. <strong>Strix</strong> <strong>occidentalis</strong> translates as<br />

“owl of <strong>the</strong> west” and <strong>lucida</strong> means “light” or<br />

“bright.” Unlike most owls, spotted owls have<br />

dark eyes. Several thin white bands mark an<br />

o<strong>the</strong>rwise brown tail.<br />

Adult male and female spotted owls are<br />

mostly monochromatic in plumage characteristics,<br />

but <strong>the</strong> sexes can be readily distinguished by<br />

voice (see below). Juveniles, subadults, and<br />

adults can be distinguished by plumage characteristics<br />

(Forsman 1981, Moen et al. 1991).<br />

Juvenile spotted owls (hatchling to approximately<br />

five months) have a downy appearance.<br />

Subadults (5 to 26 months) closely resemble<br />

adults, but have pointed retrices with a pure<br />

white terminal band (Forsman 1981, Moen et al.<br />

1991). The retrices of adults (>27 months) have<br />

rounded tips, and <strong>the</strong> terminal band is mottled<br />

brown and white.<br />

Although <strong>the</strong> spotted owl is often referred to<br />

as a medium-sized owl, it ranks among <strong>the</strong><br />

largest owls in North America. Of <strong>the</strong> 19 species<br />

of owls that occur in North America, only 4 are<br />

larger than <strong>the</strong> spotted owl (Johnsgard 1988).<br />

Like many o<strong>the</strong>r owls, spotted owls exhibit<br />

reversed sexual dimorphism (i.e., females are<br />

larger than males). Adult male <strong>Mexican</strong> spotted<br />

owls (n = 37) average 519 + 32.6 (SD) g<br />

(18.5 oz), and adult females (n = 31) average 579<br />

+ 31.2 g (20.7 oz) (Kristan et al., in prep.).<br />

There appears to be clinal variation among <strong>the</strong><br />

three subspecies in a number of morphological<br />

characteristics measured, with size decreasing<br />

from north to south (Kristan et al., in prep.).<br />

DISTRIBUTION<br />

DISTRIBUTION<br />

AND AND ABUNDANCE<br />

ABUNDANCE<br />

The <strong>Recovery</strong> Team ga<strong>the</strong>red and examined<br />

in<strong>for</strong>mation on <strong>the</strong> distribution and abundance<br />

of <strong>Mexican</strong> spotted owls through 1993. Data<br />

from surveys conducted after 1993 were not<br />

available <strong>for</strong> our analyses.<br />

21<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

We used <strong>the</strong> in<strong>for</strong>mation collected to (1)<br />

document historical and current range of this<br />

subspecies, (2) help <strong>for</strong>mulate <strong>Recovery</strong> Unit<br />

boundaries, and (3) provide a template <strong>for</strong><br />

analyses at <strong>the</strong> landscape scale. Descriptions of<br />

<strong>Recovery</strong> Units are provided in <strong>the</strong> following<br />

chapter (II.B).<br />

The <strong>Mexican</strong> spotted owl currently occupies<br />

a broad geographic area, but does not occur<br />

uni<strong>for</strong>mly throughout its range (Figure II.A.2).<br />

Instead, <strong>the</strong> owl occurs in disjunct localities that<br />

correspond to isolated mountain systems and<br />

canyons. In <strong>the</strong> United States, 91% of <strong>the</strong> owls<br />

known to exist between 1990 and 1993 occur on<br />

lands administered by <strong>the</strong> FS (Table II.A.1).<br />

O<strong>the</strong>r lands currently occupied by <strong>Mexican</strong><br />

spotted owls in <strong>the</strong> United States include, NPS<br />

(4%), BLM (2%), Tribal (2%), and DOD (1%).<br />

We know that more owls occur on Tribal lands<br />

than indicated here, but specific in<strong>for</strong>mation on<br />

numbers of owls known on Tribal lands was not<br />

made available to <strong>the</strong> Team. <strong>Owl</strong> distribution<br />

according to land ownership is unavailable <strong>for</strong><br />

Mexico. Eighty-nine percent of <strong>the</strong> owls known<br />

to exist between 1990 and 1993 in Mexico were<br />

in <strong>the</strong> States of Sonora and Chihuahua (Table<br />

II.A.1, Figure II.A.3). However, most survey<br />

ef<strong>for</strong>ts in Mexico were restricted to <strong>the</strong>se states,<br />

and <strong>the</strong>se numbers do not necessarily reflect<br />

actual trends in distribution.<br />

The current owl distribution mimics its<br />

historical extent, with a few exceptions. The owl<br />

has not been reported recently along major<br />

riparian corridors in Arizona and New Mexico,<br />

nor in historically documented areas of sou<strong>the</strong>rn<br />

Mexico. Riparian communities and previously<br />

occupied localities in <strong>the</strong> southwestern United<br />

States and sou<strong>the</strong>rn Mexico have undergone<br />

significant habitat alteration since <strong>the</strong> historical<br />

sightings (USDI 1993). However, <strong>the</strong> amount of<br />

ef<strong>for</strong>t devoted to surveying <strong>the</strong>se areas is unknown<br />

and future surveys may document<br />

spotted owls <strong>the</strong>re. Surveys conducted to relocate<br />

spotted owls in nor<strong>the</strong>rn Colorado near Fort<br />

Collins and Boulder, where records exist from<br />

<strong>the</strong> early 1970s and 1980s, have been unsuccessful.<br />

Surveys conducted in <strong>the</strong> Book Cliffs of eastcentral<br />

Utah, where owls were recorded in 1958,<br />

have also been unsuccessful. Although historical<br />

(pre-1990) data provide some in<strong>for</strong>mation about


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e II.A.2. II.A.2. Current distribution of <strong>Mexican</strong> spotted owls in <strong>the</strong> United States based on planned<br />

surveys and incidental observations recorded from 1990 through 1993.<br />

22


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able II.A.1. II.A.1. Historical records and minimum numbers of <strong>Mexican</strong> spotted owls found during<br />

planned surveys and incidental observations, reported by <strong>Recovery</strong> Unit and land ownership. <strong>Recovery</strong><br />

Units are described in Part II.B.<br />

Number umber of of 1 N NNumber<br />

N umber of<br />

of<br />

owl wl r rrecor<br />

rr<br />

ecor ecords ecor ds o oowl<br />

o wl sites<br />

sites<br />

Reco eco eco ecover eco er er ery er y U UUnit<br />

U nit be<strong>for</strong> be<strong>for</strong>e be<strong>for</strong> e 1990 1990<br />

1990-1993<br />

1990-1993<br />

UNITED UNITED ST STATES ST TES<br />

Colorado Plateau<br />

FS 21 16<br />

BLM 6 10<br />

NPS 34 23<br />

Tribal 20 13 2<br />

New Mexico State 1 0<br />

Unknown 3 5 0<br />

Subtotal ubtotal Sou<strong>the</strong>rn Rocky Mountains – Colorado<br />

87 87<br />

62<br />

62<br />

FS 2 8<br />

BLM 0 6<br />

NPS 0 0<br />

Tribal 1 -- 2<br />

Unknown 3 17 0<br />

Subtotal ubtotal Sou<strong>the</strong>rn Rocky Mountains – New Mexico<br />

20 20 20<br />

14<br />

14<br />

FS 25 34<br />

NPS 3 0<br />

New Mexico State 1 0<br />

Private 4 0<br />

Unknown 3 8 0<br />

Subtotal ubtotal Upper Gila Mountains<br />

41 41<br />

34<br />

34<br />

FS 138 424<br />

BLM 5 0<br />

NPS 5 0<br />

Tribal 0 -- 2<br />

Private 1 0<br />

Unknown 3 104 0<br />

Subtotal ubtotal Basin and Range – West<br />

253 253<br />

424<br />

424<br />

FS 82 97<br />

NPS 13 0<br />

Tribal 0 -- 2<br />

DOD 9 6<br />

Private 8 0<br />

Unknown 3 57 0<br />

Subtotal ubtotal 169 169<br />

103<br />

103<br />

23


Table able II.A.1. II.A.1. continued<br />

Number umber of of 1 Number umber of<br />

of<br />

owl wl r rrecor<br />

rr<br />

ecor ecords ecor ds owl wl sites<br />

sites<br />

Reco eco eco ecover eco er ery er y U UUnit<br />

U Unit<br />

nit be<strong>for</strong> be<strong>for</strong>e be<strong>for</strong> e 1990 1990 1990<br />

1990-1993 1990-1993<br />

1990-1993<br />

Basin and Range – East<br />

FS 18 111<br />

BLM 1 0<br />

NPS 6 10<br />

Tribal 2 -- 2<br />

FWS 1 0<br />

Private 2 0<br />

Subtotal ubtotal 30 30<br />

121<br />

121<br />

United nited nited SS<br />

States S States<br />

tates Total otal 600 600<br />

758<br />

758<br />

MEXICO<br />

MEXICO<br />

Sierra Madre Occidental – Norte<br />

Sonora 8 9<br />

Chihuahua 10 8 4<br />

Subtotal ubtotal Sierra Madre Oriental – Norte<br />

18 18<br />

17<br />

17<br />

Coahuila<br />

Sierra Madre Occidental – Sur<br />

2 0<br />

Sinaloa 1 0<br />

Durango 2 0<br />

Aguacalientes 0 1 4<br />

Zacatecas 0 0 4<br />

San Luis Potosi 1 0<br />

Guanajuato 1 0<br />

Subtotal ubtotal Sierra Madre Oriental – Sur<br />

5 5<br />

1<br />

1<br />

Coahuila 4 0<br />

Nuevo Leon 4 1<br />

Tamaulipas 0 0 4<br />

Subtotal ubtotal Eje Neovolcanico<br />

8 8<br />

1<br />

1<br />

Jalisco 1 0<br />

Colima 1 5 0<br />

Michoacan 1 0<br />

Puebla 1 5 0<br />

Subtotal ubtotal 2 2<br />

0<br />

0<br />

Mexico exico exico Total otal 35 35 35<br />

19<br />

19<br />

Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

1<br />

These values do not necessarily indicate numbers of owls or owl sites because multiple records may exist from <strong>the</strong><br />

same site through time.<br />

2 Additional owls are known to exist on many Tribal lands, but <strong>the</strong> exact number is unavailable.<br />

3<br />

Locations of <strong>the</strong>se records were insufficient <strong>for</strong> assigning a land ownership.<br />

4 Additional sightings have been reported from 1994 surveys.<br />

5 Unverified record not included in totals (see text).<br />

24


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e II.A.3. II.A.3. Current distribution of <strong>Mexican</strong> spotted owls in Mexico based on planned surveys<br />

and incidental observations recorded from 1990 through 1993.<br />

25


about past distribution of spotted owls, we stress<br />

that it is not sufficient to allow us to estimate<br />

changes in <strong>the</strong> number or distribution of owls<br />

from historical to present time.<br />

Most current observations of <strong>Mexican</strong><br />

spotted owls are from <strong>the</strong> Upper Gila Mountains<br />

RU (see II.B). This unit can be considered a<br />

critical nucleus <strong>for</strong> <strong>the</strong> subspecies because of its<br />

central location within <strong>the</strong> owl’s range and its<br />

seemingly high number of owls. O<strong>the</strong>r areas<br />

likely to be important population centers include<br />

<strong>the</strong> sky islands of sou<strong>the</strong>astern Arizona and<br />

<strong>the</strong> Sacramento Mountains of central New<br />

Mexico (Basin and Range RUs; see II.B). Although<br />

in<strong>for</strong>mation on owl numbers permits a<br />

view of <strong>the</strong> current distribution, it is not complete<br />

enough to provide a reliable estimate of<br />

total population size.<br />

<strong>Mexican</strong> spotted owls occur at higher densities<br />

in mixed-conifer <strong>for</strong>ests than in pine-oak,<br />

pine, and pinyon-juniper <strong>for</strong>est types (Skaggs<br />

and Raitt 1988, White et al. 1995). A combined<br />

estimate of <strong>Mexican</strong> spotted owl density on two<br />

study areas in <strong>the</strong> Upper Gila Mountains RU is<br />

similar to estimates reported <strong>for</strong> o<strong>the</strong>r spotted<br />

owl subspecies.<br />

In summary, <strong>the</strong> <strong>Mexican</strong> spotted owl is<br />

distributed discontinuously throughout its<br />

range, with its distribution largely restricted to<br />

montane <strong>for</strong>ests and canyons. Although future<br />

ef<strong>for</strong>ts will undoubtedly discover additional<br />

owls, <strong>the</strong>ir documented spatial distribution in<br />

<strong>the</strong> United States is not likely to change greatly.<br />

The converse is true <strong>for</strong> Mexico, where planned<br />

surveys have begun only recently.<br />

HABITAT HABITAT HABITAT ASSOCIATIONS<br />

ASSOCIATIONS<br />

<strong>Mexican</strong> spotted owls nest, roost, <strong>for</strong>age, and<br />

disperse in a diverse array of biotic communities.<br />

Mixed-conifer <strong>for</strong>ests are commonly used<br />

throughout most of <strong>the</strong> range (Johnson and<br />

Johnson 1985, Skaggs and Raitt 1988, Ganey et<br />

al. 1988, Ganey and Balda 1989a, Rinkevich<br />

1991, Willey 1993, Fletcher and Hollis 1994,<br />

Seamans and Gutiérrez, in press). In general,<br />

<strong>the</strong>se <strong>for</strong>ests are dominated by Douglas-fir and/<br />

or white fir, with codominant species including<br />

southwestern white pine, limber pine, and<br />

ponderosa pine (Brown et al. 1980). The under-<br />

Volume I/Part II<br />

26<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

story often contains <strong>the</strong> above coniferous species<br />

as well as broadleaved species such as Gambel<br />

oak, maples, boxelder, and New Mexico locust.<br />

In sou<strong>the</strong>rn Arizona and Mexico, Madrean pineoak<br />

<strong>for</strong>ests are also used commonly (Ganey and<br />

Balda 1989a, Duncan and Taiz 1992, Ganey et<br />

al. 1992, Tarango et al. 1994). These <strong>for</strong>ests are<br />

typically dominated by an overstory of Chihuahua<br />

and Apache pines in conjunction with<br />

species such as Douglas-fir, ponderosa pine, and<br />

Arizona cypress. Evergreen oaks are typically<br />

prominent in <strong>the</strong> understory (Brown et al.<br />

1980).<br />

Habitat-use patterns vary throughout <strong>the</strong><br />

range and with respect to owl activity (see<br />

below). In <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> range,<br />

including sou<strong>the</strong>rn Utah, sou<strong>the</strong>rn Colorado,<br />

and far nor<strong>the</strong>rn Arizona and New Mexico, owls<br />

occur primarily in steep-walled, rocky canyons<br />

(Kertell 1977, Reynolds 1990, Rinkevich 1991,<br />

Willey 1993). Along <strong>the</strong> Mogollon Rim in<br />

Arizona and New Mexico, habitat use is less<br />

restricted, and spotted owls occur in mixedconifer<br />

<strong>for</strong>ests, ponderosa pine-Gambel oak<br />

<strong>for</strong>ests, rocky canyons, and associated riparian<br />

<strong>for</strong>ests (Ganey and Balda 1989a, Ganey et al.<br />

1992, Fletcher and Hollis 1994, Seamans and<br />

Gutiérrez, in press, Peter Stacey, Univ. of<br />

Nevada, Reno, pers. comm.). South of <strong>the</strong><br />

Mogollon Rim and into Mexico a still wider<br />

variety of habitat types are used, including<br />

mixed-conifer, Madrean pine-oak, and Arizona<br />

cypress <strong>for</strong>ests, encinal oak woodlands, and<br />

associated riparian <strong>for</strong>ests (Ganey and Balda<br />

1989a, Duncan and Taiz 1992, Ganey et al.<br />

1992, Tarango et al. 1994). Much of this regional<br />

variation in habitat use likely results from<br />

differences in regional patterns of habitat and<br />

prey availability.<br />

Nesting Nesting and and Roosting Roosting Habitat<br />

Habitat<br />

<strong>Mexican</strong> spotted owls nest and roost primarily<br />

in closed-canopy <strong>for</strong>ests or rocky canyons. In<br />

<strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> range (sou<strong>the</strong>rn<br />

Utah and Colorado), most nests are in caves or<br />

on cliff ledges in steep-walled canyons. Elsewhere,<br />

<strong>the</strong> majority of nests appear to be in trees<br />

(Fletcher and Hollis 1994).


Forests used <strong>for</strong> roosting and nesting often<br />

contain mature or old-growth stands with<br />

complex structure (Skaggs and Raitt 1988,<br />

Ganey and Balda 1989a, 1994; McDonald et al.<br />

1991, Seamans and Gutiérrez, in press). These<br />

<strong>for</strong>ests are typically uneven-aged, multistoried,<br />

and have high canopy closure. Nest trees are<br />

typically large in size (SWCA 1992, Fletcher and<br />

Hollis 1994, Seamans and Gutiérrez, in press),<br />

whereas owls roost in both large and small trees<br />

(Ganey 1988, Rinkevich 1991, Willey 1993,<br />

Zwank et al. 1994, Peter Stacey, Univ. of<br />

Nevada, Reno, pers. comm.). Tree species used<br />

<strong>for</strong> nesting vary somewhat among areas and<br />

habitat types, but available evidence suggests that<br />

Douglas-fir is <strong>the</strong> most common species of nest<br />

tree (SWCA 1992, Fletcher and Hollis 1994,<br />

Seamans and Gutiérrez, in press). A wider variety<br />

of trees are used <strong>for</strong> roosting, but again Douglasfir<br />

is <strong>the</strong> most commonly used species (Ganey<br />

1988, Fletcher and Hollis 1994, Zwank et al.<br />

1994, Peter Stacey, Univ. of Nevada, Reno, pers.<br />

comm.).<br />

Several hypo<strong>the</strong>ses have been proposed to<br />

explain why spotted owls nest in closed-canopy<br />

<strong>for</strong>ests (reviewed by Carey 1985, Gutiérrez<br />

1985). Barrows (1981) suggested that spotted<br />

owls are relatively intolerant of high temperatures<br />

and roost and nest in shady <strong>for</strong>ests because<br />

<strong>the</strong>y provide favorable microclimatic conditions.<br />

Ganey et al. (1993) observed that <strong>Mexican</strong><br />

spotted owls produced more metabolic heat than<br />

great horned owls, and were less able to dissipate<br />

that heat. This may lead <strong>the</strong>se owls to seek out<br />

cool microsites during periods of high ambient<br />

temperature. <strong>Mexican</strong> spotted owls typically nest<br />

and roost in closed-canopy <strong>for</strong>ests or deep shady<br />

canyons; both situations provide cool microsites<br />

(Kertell 1977, Ganey et al. 1988, Rinkevich<br />

1991, Ganey and Balda 1989a, Willey 1993).<br />

Foraging Foraging Habitat<br />

Habitat<br />

Little is known about patterns of habitat use<br />

by <strong>for</strong>aging owls. The only available data describe<br />

habitat use by eight owls occupying five<br />

home ranges on three study areas in nor<strong>the</strong>rn<br />

Arizona (Ganey and Balda 1994). In general,<br />

owls <strong>for</strong>aged more than or as expected in<br />

unlogged <strong>for</strong>ests, and less than or as expected in<br />

Volume I/Part II<br />

27<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

selectively logged <strong>for</strong>ests. Expected values were<br />

based on relative occurrences of habitats. However,<br />

patterns of habitat use varied among study<br />

areas and individuals, and even between pair<br />

members in some cases, making generalizations<br />

difficult. Both high-use roosting and high-use<br />

<strong>for</strong>aging sites had more big logs, higher canopy<br />

closure, and greater densities and basal areas of<br />

both trees and snags than random sites. <strong>Owl</strong>s<br />

clearly used a wider variety of <strong>for</strong>est conditions<br />

<strong>for</strong> <strong>for</strong>aging than <strong>the</strong>y used <strong>for</strong> roosting (Ganey<br />

and Balda 1994). We caution, however, about<br />

extending <strong>the</strong>se results too far given <strong>the</strong> limited<br />

number of owls sampled, and <strong>the</strong> variability<br />

observed among owls and sites.<br />

TERRITORIALITY<br />

TERRITORIALITY<br />

AND AND HOME HOME HOME RANGE RANGE<br />

RANGE<br />

Home range is defined as <strong>the</strong> area used by an<br />

animal during its normal activities (Burt 1943)<br />

whereas territory is a defended area within an<br />

individual’s home range (Nice 1941). Territories<br />

are typically smaller than home ranges, but <strong>the</strong><br />

exact relationship between <strong>the</strong> territory and <strong>the</strong><br />

home range is generally not known. Fidelity to<br />

territories is apparently high in <strong>Mexican</strong> spotted<br />

owls, with most owls remaining on <strong>the</strong> same<br />

territory year after year.<br />

Home-range size of <strong>Mexican</strong> spotted owls,<br />

as estimated by monitoring movements of<br />

radiotagged owls, appears to vary considerably<br />

among habitats and/or geographic areas (Ganey<br />

and Dick 1995). Differences in sampling methods<br />

among studies make comparisons difficult,<br />

however. Minimum convex polygon home<br />

range estimates of home range-size varied<br />

from (1) 924 - 1,487 ha (2,282 - 3,672 acres)<br />

<strong>for</strong> individuals (n = 11) on three study areas in<br />

Colorado Plateau RU (Willey 1993); (2) 261 -<br />

1,053 ha (645 - 2,601 acres) <strong>for</strong> individuals<br />

(n = 25) and 381 - 1,551 ha (941 - 3,831 acres)<br />

<strong>for</strong> pairs (n = 10) on five study areas in <strong>the</strong><br />

Upper Gila Mountains RU (Ganey and Balda<br />

1989b, Ganey and Block, unpublished data,<br />

Peter Stacey, Univ. of Nevada, Reno, pers.<br />

comm.); and (3) 452 - 937 ha (1,116 - 2,314<br />

acres) <strong>for</strong> individuals (n = 20) and 573 - 1,401<br />

ha (1,415 - 3,461 acres) <strong>for</strong> pairs (n = 8) on two<br />

study areas in <strong>the</strong> Basin and Range-East RU


(Zwank et al. 1994, Ganey and Block, unpublished<br />

data).<br />

Volume I/Part II<br />

VOCALIZATIONS<br />

VOCALIZATIONS<br />

The spotted owl, being primarily nocturnal,<br />

is more often heard than seen. It has a wide<br />

repertoire of calls (Forsman et al. 1984, Ganey<br />

1990) that are relatively low-pitched and composed<br />

of pure tones (Fitton 1991). Sexes can be<br />

distinguished by calls; males have a deeper voice<br />

than females and generally call more frequently.<br />

The most common vocalization, used more<br />

often by males, is a series of four unevenlyspaced<br />

hoots. Females frequently use a clear<br />

whistle ending with an upward inflection as well<br />

as a series of sharp barks. Forsman et al. (1984)<br />

described 14 calls <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted owl,<br />

of which 10 were reported by Ganey (1990) <strong>for</strong><br />

Arizona birds.<br />

<strong>Mexican</strong> spotted owls call mainly from<br />

March - November and are relatively silent from<br />

December - February (Ganey 1990). Calling<br />

activity increases from March through May<br />

(although nesting females are largely silent<br />

during April and early May), and <strong>the</strong>n declines<br />

from June through November (Ganey 1990).<br />

On a daily basis, calling activity is greatest<br />

during <strong>the</strong> 2-hour period following sunset, with<br />

smaller peaks 4-8 hours after sunset and just<br />

be<strong>for</strong>e sunrise. <strong>Owl</strong>s called more than expected<br />

during <strong>the</strong> last quarter and new moon phases of<br />

<strong>the</strong> lunar cycle; and <strong>the</strong>y called most frequently<br />

on calm, clear nights when no precipitation was<br />

falling (Ganey 1990).<br />

INTERSPECIFIC INTERSPECIFIC COMPETITION<br />

COMPETITION<br />

Several o<strong>the</strong>r species of owls occur within <strong>the</strong><br />

range of <strong>the</strong> <strong>Mexican</strong> spotted owl. Interference<br />

competition, where individuals physically<br />

interfere with each o<strong>the</strong>r, probably does not<br />

occur to any great extent between <strong>the</strong> <strong>Mexican</strong><br />

spotted owl and o<strong>the</strong>r owl species. However,<br />

exploitative competition, where individuals<br />

compete <strong>for</strong> similar resources such as prey or<br />

nest sites, may occur. Competition might be<br />

greatest between spotted and great horned owls<br />

because both species are relatively large and<br />

widely distributed. Preliminary data from a<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

telemetry study in nor<strong>the</strong>rn Arizona suggest that<br />

<strong>the</strong>se owls overlap broadly in diet and space, but<br />

exhibit some habitat segregation (Ganey and<br />

Block, unpublished data). If <strong>Mexican</strong> spotted<br />

and barred owls are sympatric in Mexico, <strong>the</strong>n<br />

competition might also occur between <strong>the</strong>se<br />

closely related species. In general, however, more<br />

research is needed to assess <strong>the</strong> potential occurrence<br />

and importance of interspecific competition<br />

between spotted and o<strong>the</strong>r owls.<br />

FEEDING FEEDING HABITS HABITS AND AND PREY<br />

PREY<br />

ECOLOGY<br />

ECOLOGY<br />

Forsman (1976) described spotted owls as<br />

“perch and pounce” predators. They typically<br />

locate prey from an elevated perch by sight or<br />

sound, <strong>the</strong>n pounce on <strong>the</strong> prey and capture it<br />

with <strong>the</strong>ir talons. <strong>Spotted</strong> owls have also been<br />

observed capturing flying prey such as birds and<br />

insects (Verner et al. 1992b). They hunt primarily<br />

at night (Forsman et al. 1984, Ganey 1988),<br />

although infrequent diurnal <strong>for</strong>aging has been<br />

documented (Forsman et al. 1984, Laymon<br />

1991, Sovern et al. 1994).<br />

<strong>Mexican</strong> spotted owls consume a variety of<br />

prey throughout <strong>the</strong>ir range but commonly eat<br />

small- and medium-sized rodents such as<br />

woodrats, peromyscid mice, and microtine voles.<br />

<strong>Spotted</strong> owls also consume bats, birds, reptiles,<br />

and arthropods. The diet varies by geographic<br />

location (Ward and Block 1995; Figure. II.A.4).<br />

For example, spotted owls dwelling in canyons<br />

of <strong>the</strong> Colorado Plateau take more woodrats,<br />

and fewer birds, than do spotted owls from o<strong>the</strong>r<br />

areas (Ward and Block 1995, Figure. II.A.4). In<br />

contrast, spotted owls occupying mountain<br />

ranges with <strong>for</strong>est-meadow interfaces, as found<br />

within <strong>the</strong> Basin and Range - East, Sou<strong>the</strong>rn<br />

Rocky Mountains - Colorado, and Upper Gila<br />

Mountains RUs, take more voles (Ward and<br />

Block 1995, Figure. II.A.4). Regional differences<br />

in <strong>the</strong> owl’s diet likely reflect geographic variation<br />

in population densities and habitats of both<br />

<strong>the</strong> prey and <strong>the</strong> owl.<br />

The Team was unable to link consumption<br />

of specific prey and successful reproduction by<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl, with two possible<br />

exceptions. First, fecundity of spotted owls<br />

occupying <strong>the</strong> Sacramento Mountains (Basin<br />

28


and Range - East RU) appeared to be associated<br />

with trends in abundance of peromyscid mice<br />

(Ward and Block 1995). Second, <strong>the</strong> predominance<br />

of woodrats in <strong>the</strong> diet throughout much<br />

of <strong>the</strong> owl’s range suggests that this prey may<br />

influence <strong>the</strong> owl’s fitness. O<strong>the</strong>r studies have<br />

shown positive associations between larger prey<br />

(e.g., woodrats) in <strong>the</strong> diet of nor<strong>the</strong>rn and<br />

Cali<strong>for</strong>nia spotted owls and <strong>the</strong>ir reproductive<br />

success (Barrows 1987, Thrailkill and Bias<br />

1989). In most cases, however, total prey biomass<br />

may be more influential on <strong>the</strong> owl’s fitness<br />

than <strong>the</strong> abundance of any particular prey<br />

species.<br />

Habitat correlates of <strong>the</strong> owl’s common prey<br />

emphasize that each prey species uses a unique<br />

microhabitat. For example, deer mice are ubiquitous<br />

in distribution, occupying areas with<br />

variable conditions, whereas brush mice are<br />

restricted to communities with a strong oak<br />

component and dry, rocky substrates with sparse<br />

tree cover. <strong>Mexican</strong> woodrats are typically found<br />

in areas with considerable shrub or understory<br />

tree cover, little herbaceous cover, and high log<br />

volumes. <strong>Mexican</strong> voles are found in areas with<br />

high herbaceous cover, primarily grasses. Longtailed<br />

voles are associated with high herbaceous<br />

cover, primarily <strong>for</strong>bs, many shrubs, and limited<br />

tree cover. Thus, to provide a diverse prey base,<br />

managers should provide diverse habitats <strong>for</strong><br />

prey species. Managing habitat <strong>for</strong> a diversity of<br />

prey species may help buffer against population<br />

fluctuations of individual prey species and<br />

provide a more constant food supply <strong>for</strong> <strong>the</strong> owl.<br />

REPRODUCTIVE REPRODUCTIVE BIOLOGY<br />

BIOLOGY<br />

Knowledge of <strong>the</strong> annual reproductive cycle<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl is important both in<br />

an ecological context, and <strong>for</strong> placing seasonal<br />

restrictions on management or on o<strong>the</strong>r activities<br />

that may occur within areas occupied by spotted<br />

owls. Data on <strong>the</strong> reproductive cycle of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl are limited compared to<br />

in<strong>for</strong>mation on <strong>the</strong> nor<strong>the</strong>rn and Cali<strong>for</strong>nia<br />

subspecies. There<strong>for</strong>e, although <strong>the</strong> following<br />

discussion is based primarily on observations of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl, data from <strong>the</strong> o<strong>the</strong>r<br />

subspecies are provided to fill some in<strong>for</strong>mation<br />

gaps.<br />

Volume I/Part II<br />

29<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>Mexican</strong> spotted owls nest on cliff ledges,<br />

stick nests built by o<strong>the</strong>r birds, debris plat<strong>for</strong>ms<br />

in trees, and in tree cavities (Johnson and<br />

Johnson 1985, Ganey 1988, SWCA 1992,<br />

Fletcher and Hollis 1994, Seamans and<br />

Gutiérrez, in press). <strong>Spotted</strong> owls have one of<br />

<strong>the</strong> lowest clutch sizes among North American<br />

owls (Johnsgard 1988). Females normally lay<br />

one to three eggs, two being most common. Renesting<br />

following nest failure is unusual, but has<br />

been observed in <strong>Mexican</strong> spotted owls (Kroel<br />

1991, David Olson, Humboldt State Univ.,<br />

Arcata, CA, pers. comm.). <strong>Mexican</strong> spotted owls<br />

breed sporadically and do not nest every year<br />

(Ganey 1988). In good years most of <strong>the</strong> population<br />

will nest, whereas in o<strong>the</strong>r years only a<br />

small proportion of pairs will nest successfully<br />

(Fletcher and Hollis 1994). Reasons <strong>for</strong> this<br />

pattern are unknown.<br />

<strong>Mexican</strong> spotted owls have distinct<br />

annual breeding periods, but reproductive<br />

chronology varies somewhat across <strong>the</strong> range of<br />

<strong>the</strong> owl. In Arizona, courtship apparently begins<br />

in March with pairs roosting toge<strong>the</strong>r during <strong>the</strong><br />

day and calling to each o<strong>the</strong>r at dusk (Ganey<br />

1988). Eggs are laid in late March or, more<br />

typically, early April. Incubation begins shortly<br />

after <strong>the</strong> first egg is laid, and is per<strong>for</strong>med<br />

entirely by <strong>the</strong> female (Ganey 1988). Female<br />

nor<strong>the</strong>rn spotted owls incubate <strong>for</strong> approximately<br />

30 days (Forsman et al. 1984), and<br />

<strong>Mexican</strong> spotted owls appear to incubate <strong>for</strong> a<br />

similar period (Ganey 1988). During incubation<br />

and <strong>the</strong> first half of <strong>the</strong> brooding period, <strong>the</strong><br />

female leaves <strong>the</strong> nest only to defecate, regurgitate<br />

pellets, or to receive prey delivered by <strong>the</strong><br />

male, who does most or all of <strong>the</strong> <strong>for</strong>aging<br />

(Forsman et al. 1984, Ganey 1988).<br />

The eggs usually hatch in early May (Ganey<br />

1988). Females brood <strong>the</strong>ir young almost<br />

constantly <strong>for</strong> <strong>the</strong> first couple of weeks after <strong>the</strong><br />

eggs hatch but <strong>the</strong>n begin to spend time hunting<br />

at night, leaving <strong>the</strong> owlets unattended <strong>for</strong> up to<br />

several hours (Eric Forsman, FS, Corvallis, OR,<br />

pers. comm.). Nestling owls generally fledge four<br />

to five weeks after hatching in early to mid-June<br />

(Ganey 1988). <strong>Owl</strong>ets usually leave <strong>the</strong> nest<br />

be<strong>for</strong>e <strong>the</strong>y can fly, jumping from <strong>the</strong> nest on to<br />

surrounding tree branches or <strong>the</strong> ground<br />

(Forsman et al. 1984, Ganey 1988). <strong>Owl</strong>ets that


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igure<br />

e II.A.4. II.A.4. Geographic variability in <strong>the</strong> food habits of <strong>Mexican</strong> spotted owls presented as relative<br />

frequencies of (a) (a) (a) woodrats, (b) (b) voles, (c) (c) gophers, (d) (d) birds, (e) (e) bats, and (f (f (f) (f (f reptiles. Point values are<br />

from single studies or averages among <strong>the</strong> number of studies shown in paren<strong>the</strong>sis (a) (a) (a). (a) (a) Vertical bars are<br />

95% confidence intervals showing sampling and interstudy variation within a recovery unit. <strong>Recovery</strong><br />

unit acronyms are COPLAT-Colorado Plateau; SRM-CO-Sou<strong>the</strong>rn Rocky Mountain-Colorado; SRM-<br />

NM-Sou<strong>the</strong>rn Rocky Mountain-New Mexico; UPGIL-Upper Gila Mountain; BAR-W-Basin and<br />

Range - West; BAR-E-Basin and Range - East; SMO-N-Sierra Madre Occidental-Norte.<br />

30


end up on <strong>the</strong> ground will often climb back up a<br />

tree to a safe roost site. The mobility and <strong>for</strong>aging<br />

skills of owlets improve gradually during <strong>the</strong><br />

summer. Within a week after leaving <strong>the</strong> nest,<br />

most owlets can make short, clumsy flights<br />

between trees. Three weeks after leaving <strong>the</strong> nest,<br />

owlets can hold and tear up prey on <strong>the</strong>ir own<br />

(Forsman et al. 1984).<br />

Fledglings depend on <strong>the</strong>ir parents <strong>for</strong> food<br />

during <strong>the</strong> early portion of <strong>the</strong> fledgling period.<br />

Hungry owlets give a persistent, raspy “begging<br />

call,” especially when adults appear with food or<br />

call nearby (Forsman et al. 1984, Ganey 1988).<br />

Begging behavior declines in late August, but<br />

may continue at low levels until dispersal occurs,<br />

usually from mid September to early October<br />

(Ganey and Block, unpubl. data, Peter Stacey,<br />

Univ. of Nevada, Reno, pers. comm., David<br />

Willey, Nor<strong>the</strong>rn Arizona Univ., Flagstaff, pers.<br />

comm.).<br />

Volume I/Part II<br />

MORTALITY MORTALITY MORTALITY FACTORS<br />

FACTORS<br />

Several mortality factors (discussed below)<br />

have been identified as potentially important<br />

with respect to <strong>the</strong> <strong>Mexican</strong> spotted owl. Although<br />

a number of owls have been recovered<br />

following mortality and examined by both field<br />

biologists and laboratory personnel, in general<br />

little is known about <strong>the</strong> extent or importance of<br />

<strong>the</strong>se mortality factors.<br />

Predation<br />

Predation<br />

Predation, particularly by avian predators,<br />

may be a common mortality factor of spotted<br />

owls. Potential avian predators of <strong>Mexican</strong><br />

spotted owls include great horned owls, nor<strong>the</strong>rn<br />

goshawks, red-tailed hawks, and golden eagles.<br />

Some of <strong>the</strong>se predators occupy <strong>the</strong> same general<br />

habitats as <strong>the</strong> spotted owl, but <strong>the</strong>re is little<br />

direct evidence that <strong>the</strong>y prey on spotted owls to<br />

any great extent (Gutiérrez et al. 1995). Ganey<br />

(1988) reported one instance of apparent great<br />

horned owl predation on an adult spotted owl,<br />

and Richard Reynolds (FS, Fort Collins, CO.,<br />

pers. comm.) reported a golden eagle preying on<br />

a spotted owl. Preliminary results from radiotagged<br />

<strong>Mexican</strong> spotted owls indicate that both<br />

31<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

adults and juveniles are preyed upon (Willey<br />

1993, Ganey and Block, unpubl. data), but in<br />

most cases <strong>the</strong> identity of <strong>the</strong> predator was<br />

unknown. Fur<strong>the</strong>r, in sou<strong>the</strong>rn Arizona,<br />

procyonid mammals were observed attempting<br />

to raid cliff site nests occupied by spotted owls<br />

(Russell Duncan, Southwestern Field Biologists,<br />

Tucson, AZ, pers. comm.). Thus, <strong>the</strong> extent to<br />

which <strong>Mexican</strong> spotted owls are preyed upon is<br />

unknown at this time.<br />

Starvation<br />

Starvation<br />

Starvation is likely ano<strong>the</strong>r common source<br />

of mortality. Juvenile nor<strong>the</strong>rn spotted owls may<br />

be more vulnerable to starvation than adults<br />

(Gutiérrez et al. 1985, Miller 1989), because of<br />

<strong>the</strong>ir poor hunting skills. Starvation may also<br />

result from low abundance or availability of prey,<br />

which could affect both adults and juveniles.<br />

Both adult and juvenile owls radio-tagged in<br />

Arizona have been found dead of apparent<br />

starvation (Ganey and Block, unpublished data),<br />

and two of seven radio-tagged juveniles in Utah<br />

died of starvation (Willey 1993). Most instances<br />

of starvation occurred from late fall through<br />

winter, when prey resources were reduced in<br />

abundance and availability (Willey 1993, Block<br />

and Ganey, unpublished data). In addition,<br />

starvation may predispose young or even adults<br />

to predation.<br />

Accidents<br />

Accidents<br />

Accidents may be ano<strong>the</strong>r mortality factor.<br />

For example, instances of spotted owls being hit<br />

by cars have been documented (Roger Skaggs,<br />

New Mexico State Univ, Las Cruces, pers.<br />

comm; Russell Duncan, Southwestern Field<br />

Biologists, Tucson, AZ, pers. comm.). <strong>Owl</strong>s<br />

flying at night might also collide with<br />

powerlines, tree branches, or o<strong>the</strong>r obstacles.<br />

This might be particularly true <strong>for</strong> birds migrating<br />

or dispersing through unfamilar terrain.<br />

Again, little in<strong>for</strong>mation is available on how<br />

frequently this might occur.


Disease Disease and and Parasites<br />

Parasites<br />

Little is known about how disease and<br />

parasites contribute to mortality of spotted owls.<br />

Hunter et al. (1994) found a larval mite and lice<br />

on 2 of 28 museum specimens of <strong>Mexican</strong><br />

spotted owls examined <strong>for</strong> parasites, and 6 of 18<br />

live owls examined had hippoboscid fly larvae in<br />

<strong>the</strong>ir ears. Some of <strong>the</strong> live owls examined also<br />

had lice. Hunter et al. (1994) reached no conclusions<br />

concerning mortality and ectoparasites in<br />

spotted owls, but did suggest that larval infestations<br />

in <strong>the</strong>ir ears could affect <strong>the</strong> owls’ hearing.<br />

Because hearing is important <strong>for</strong> <strong>for</strong>aging at<br />

night, such infestations could impact <strong>the</strong> birds’<br />

ability to hunt effectively.<br />

In general, however, spotted owls may be<br />

adapted to high parasite loads. In a survey of<br />

blood parasites in all three subspecies of spotted<br />

owls, Gutiérrez (1989) found an infection rate of<br />

100 percent. Although disease and parasites<br />

could predispose owls to death by starvation,<br />

predation, or accident, no evidence exists documenting<br />

disease and parasites as direct mortality<br />

factors within <strong>the</strong> <strong>Mexican</strong> subspecies.<br />

Volume I/Part II<br />

POPULATION POPULATION BIOLOGY BIOLOGY<br />

BIOLOGY<br />

The <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> population <strong>for</strong> a<br />

specific area can be modeled with <strong>the</strong> simple<br />

equation<br />

N t+1 = N t + B t - D t + I t - E t ,<br />

where N t is <strong>the</strong> population size at time t, B t is <strong>the</strong><br />

number of new birds recruited into <strong>the</strong> population<br />

(births), D t is <strong>the</strong> number of birds dying, I t<br />

is <strong>the</strong> number of birds immigrating into <strong>the</strong><br />

population, and E t is <strong>the</strong> number of birds emigrating<br />

from <strong>the</strong> population. The combined<br />

effect of births, deaths, immigrations, and<br />

emigrations dictate <strong>the</strong> viability of <strong>the</strong> population,<br />

and hence its long-term persistence.<br />

Survival<br />

Survival<br />

Annual survival rates of adult <strong>Mexican</strong><br />

spotted owls is 0.8-0.9 based on short-term<br />

population and radio-tracking studies and<br />

longer-term monitoring studies (White et al.<br />

32<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

1995). These annual survival estimates can be<br />

viewed as <strong>the</strong> probability of an individual surviving<br />

from one year to <strong>the</strong> next or as <strong>the</strong> proportion<br />

of individuals that will survive from one<br />

year to <strong>the</strong> next. A variety of different estimators<br />

of adult survival using different types and sets of<br />

data gave similar results.<br />

Juvenile survival is considerably lower ( 0.06-<br />

0.29) than adult survival. Juvenile survival also<br />

appears more spatially variable, although this<br />

conclusion reflects only two population study<br />

areas and two radio-telemetry studies spanning<br />

two years or less.<br />

We strongly suspect that estimates of juvenile<br />

survival from <strong>the</strong> population studies which<br />

utilize mark-recapture methods are biased low<br />

because of (1) a high likelihood of permanent<br />

dispersal (emigration) from <strong>the</strong> study area, and<br />

(2) a lag of several years be<strong>for</strong>e marked juveniles<br />

reappear as territory holders, at which point <strong>the</strong>y<br />

are first detected <strong>for</strong> recapture. Juvenile nor<strong>the</strong>rn<br />

spotted owls have a high dispersal capability<br />

(reviewed in Thomas et al. 1990). If <strong>Mexican</strong><br />

spotted owl juveniles have a similar dispersal<br />

capability, we expect that a substantial portion of<br />

marked juveniles will emigrate from <strong>the</strong> respective<br />

study areas. However, estimates from <strong>the</strong><br />

radio-telemetry studies roughly corroborated <strong>the</strong><br />

low estimates from <strong>the</strong> population studies. Biases<br />

in <strong>the</strong> radio-telemetry estimates of juvenile<br />

survival can result if radios significantly affect<br />

<strong>the</strong>ir survival. Whe<strong>the</strong>r radios or <strong>the</strong>ir attachment<br />

affect survival of nor<strong>the</strong>rn spotted owls is<br />

debatable (Paton et al. 1991, Foster et al. 1992).<br />

Concerning <strong>the</strong> second point, Franklin (1992)<br />

found a lag of 1-4 years between <strong>the</strong> time when<br />

juvenile nor<strong>the</strong>rn spotted owls were banded and<br />

subsequently recaptured. If this process is similar<br />

<strong>for</strong> <strong>Mexican</strong> spotted owls, <strong>the</strong>n <strong>the</strong> current<br />

population studies may be of insufficient duration<br />

to adequately estimate juvenile survival.<br />

In summary, current survival estimates are<br />

based primarily on studies of insufficient duration<br />

or studies not explicitly designed to estimate<br />

survival. In most cases, <strong>the</strong> data are too limited<br />

to support or test <strong>the</strong> assumptions of <strong>the</strong> estimators<br />

used. However, <strong>the</strong> age- and sex-specific<br />

estimates of survival calculated here are useful at<br />

this point as qualitative descriptors of <strong>the</strong> lifehistory<br />

characteristics of <strong>Mexican</strong> spotted owls.


That is, <strong>Mexican</strong> spotted owls exhibit high adult<br />

and relatively low juvenile survival. In this<br />

respect, <strong>Mexican</strong> spotted owl survival probabilities<br />

appear similar to nor<strong>the</strong>rn (see review in<br />

Burnham et al. 1994) and Cali<strong>for</strong>nia spotted<br />

owls (Noon et al. 1992).<br />

Reproduction<br />

Reproduction<br />

Reproductive output of <strong>Mexican</strong> spotted<br />

owls, defined as <strong>the</strong> number of young fledged<br />

per pair, varies both spatially and temporally<br />

(White et al. 1995). <strong>Mexican</strong> spotted owls may<br />

have a higher average reproductive rate (1.001<br />

fledged young per pair) than <strong>the</strong> Cali<strong>for</strong>nia<br />

(~0.712; Noon et al. 1992) and <strong>the</strong> nor<strong>the</strong>rn<br />

spotted owl (~0.715; Thomas et al. 1993). All<br />

three subspecies exhibit temporal fluctuations in<br />

reproduction, although <strong>the</strong> amplitude of those<br />

fluctuations may be greatest <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl.<br />

Environmental Environmental Variation<br />

Variation<br />

Environmental conditions greatly affect<br />

reproduction and/or survival of nestlings<br />

through fledging and to adulthood. However,<br />

adult survival rates appear to be relatively constant<br />

across years, as suggested by high pair<br />

persistence rates (White et al. 1995). Such life<br />

history characteristics are common <strong>for</strong> Kselected<br />

species, <strong>for</strong> which populations remain<br />

relatively stable even though recruitment rates<br />

might be highly variable. With no recruitment,<br />

<strong>the</strong> population declines at <strong>the</strong> rate of 1 minus<br />

adult survival, or <strong>the</strong> adult mortality rate.<br />

Population Population Trends Trends<br />

Trends<br />

We have inadequate data to estimate population<br />

trends in <strong>Mexican</strong> spotted owls. We have<br />

little confidence in our estimates of population<br />

trend that include estimates of juvenile survival<br />

because <strong>the</strong>se estimates of juvenile survival are<br />

probably biased low. Fur<strong>the</strong>r, <strong>the</strong> population<br />

studies from which parameter estimates were<br />

derived have not been conducted <strong>for</strong> a sufficiently<br />

long period to capture temporal variation.<br />

Population trend was also evaluated with<br />

Volume I/Part II<br />

33<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

occupancy data (White et al. 1995), but again is<br />

suspect. Changes in occupancy rate probably<br />

correspond more with how monitoring of owls<br />

was per<strong>for</strong>med ra<strong>the</strong>r than reflecting true change<br />

in <strong>the</strong> owl population. As a complicating factor,<br />

a nonrandom sample of all existing <strong>Mexican</strong><br />

spotted owl territories was monitored, thus<br />

limiting possible inferences.<br />

MOVEMENTS<br />

MOVEMENTS<br />

MOVEMENTS<br />

Seasonal Seasonal Movements Movements<br />

Movements<br />

Seasonal movement patterns of <strong>Mexican</strong><br />

spotted owls are variable. Some radio-tracked<br />

owls are year-round residents within an area,<br />

some remain in <strong>the</strong> same general area but show<br />

shifts in habitat-use patterns, and some migrate<br />

considerable distances (20-50 km [12-31 miles])<br />

during <strong>the</strong> winter (Ganey and Balda 1989b,<br />

Ganey et al. 1992, Willey 1993, Ganey and<br />

Block unpublished data). In general, migrating<br />

owls move to more open habitats at lower<br />

elevations (Ganey et al. 1992, Willey 1993).<br />

Willey (1993), however, observed one owl that<br />

migrated to coniferous <strong>for</strong>est at a higher elevation<br />

than <strong>the</strong> owls’ breeding-season range.<br />

Natal Natal Dispersal<br />

Dispersal<br />

Little is known about habitat use by<br />

juveniles during natal dispersal. Seven juveniles<br />

radio-tracked in sou<strong>the</strong>rn Utah (Willey 1993)<br />

dispersed over distances ranging from 24 to 145<br />

km (15 to 90 miles). These owls apparently<br />

moved through a variety of habitats including<br />

spruce-fir and mixed-conifer <strong>for</strong>ests, pinyonjuniper<br />

woodland, mountain shrublands, desert<br />

scrublands and desert grasslands. Ano<strong>the</strong>r five<br />

juvenile owls were radio-tagged in <strong>the</strong> San Mateo<br />

Mountains of New Mexico in 1993 (Peter<br />

Stacey, Univ. of Nevada, Reno, pers. comm.).<br />

Two of <strong>the</strong>se apparently moved to an adjacent<br />

mountain range be<strong>for</strong>e <strong>the</strong>ir signals were lost. Of<br />

<strong>the</strong> remaining three, one was relocated <strong>the</strong><br />

following year within <strong>the</strong> San Mateo Mountains.<br />

Fates of <strong>the</strong> o<strong>the</strong>r two juveniles were unknown.


Volume I/Part II<br />

LANDSCAPE LANDSCAPE LANDSCAPE PATTERN<br />

PATTERN<br />

AND AND METAPOPULATION<br />

METAPOPULATION<br />

STRUCTURE<br />

STRUCTURE<br />

Keitt et al. (1995) examined <strong>the</strong> spatial<br />

pattern of <strong>for</strong>est habitat patches across <strong>the</strong> range<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl. Their objective was<br />

to gauge <strong>the</strong> extent to which <strong>the</strong> owl might<br />

behave as a metapopulation in <strong>the</strong> classical sense<br />

of a set of local populations linked by infrequent<br />

dispersal. Such a finding, if verified, would<br />

suggest that population dynamics of owls in one<br />

local population might be influenced by factors,<br />

including management activities, that affected<br />

nearby populations. Conversely, if local populations<br />

are functionally discrete, <strong>the</strong>n those populations<br />

could be treated separately with some<br />

confidence that actions in one part of <strong>the</strong> owl’s<br />

range would not greatly affect o<strong>the</strong>r populations.<br />

Keitt et al. (1995) concluded that <strong>the</strong> owl<br />

probably behaves as a classical metapopulation<br />

over much of its range. That is, <strong>the</strong> level of<br />

habitat connectivity is such that many habitats<br />

are “nearly connected” at distances corresponding<br />

to <strong>the</strong>ir best empirical estimates of <strong>the</strong> owl’s<br />

dispersal capability. At this scale, <strong>the</strong> landscape<br />

consists of a set of large, more-or-less discrete<br />

habitat clusters. For example, most of <strong>the</strong><br />

Mogollon Rim functions as a single cluster, <strong>the</strong><br />

sou<strong>the</strong>rn Rockies as ano<strong>the</strong>r single cluster, and<br />

so on. This suggests that owls could disperse<br />

within habitat clusters with very high probability,<br />

and disperse between clusters at very low<br />

probability. Thus, we would expect owls to<br />

disperse within clusters most of <strong>the</strong> time but<br />

between clusters only rarely which is consistent<br />

with <strong>the</strong> definition of a metapopulation. This<br />

finding suggests that <strong>the</strong> <strong>Plan</strong> should incorporate<br />

recommendations that maintain (or increase)<br />

habitat connectivity across <strong>the</strong> owl’s range.<br />

Habitat connectivity buffers a population from<br />

stochastic variability through time by providing<br />

<strong>the</strong> opportunity <strong>for</strong> local population failures to<br />

be “rescued” by immigration from o<strong>the</strong>r populations.<br />

Keitt et al. (1995) also attempted to identify<br />

those habitat clusters most important to overall<br />

landscape connectivity. They first ranked habitats<br />

to emphasize <strong>the</strong> importance of large patches<br />

34<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

in <strong>the</strong> landscape, and second, <strong>the</strong>y modified this<br />

approach to emphasize positional effects (i.e.,<br />

small clusters that are important because <strong>the</strong>y act<br />

as “stepping stones” or bridges between larger<br />

habitat clusters).<br />

In <strong>the</strong> first analysis, <strong>the</strong> largely contiguous<br />

habitat of <strong>the</strong> Mogollon Rim emerged as most<br />

important overall, because of its large area. In <strong>the</strong><br />

analysis emphasizing cluster position, a few small<br />

clusters emerged as particularly important. These<br />

included several fragments of <strong>the</strong> Cibola National<br />

Forest (Mt. Taylor and Zuni Mountains)<br />

that may serve as stepping stones between o<strong>the</strong>r,<br />

larger clusters. These small patches may warrant<br />

particular management attention; <strong>the</strong>y may<br />

support few owls but may never<strong>the</strong>less be<br />

important to overall landscape connectivity.<br />

CONCLUSIONS<br />

CONCLUSIONS<br />

In many ways, <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

appears to be quite similar to both <strong>the</strong> nor<strong>the</strong>rn<br />

and Cali<strong>for</strong>nia spotted owls with respect to<br />

general behavioral patterns and ecology. For<br />

example, all three subspecies are most common<br />

in <strong>for</strong>ests of complex structure, prey mainly on<br />

nocturnally-active small mammals, and share<br />

similar vocalizations, reproductive chronologies,<br />

and population characteristics. However, important<br />

differences exist between <strong>the</strong> <strong>Mexican</strong><br />

spotted owl and <strong>the</strong> o<strong>the</strong>r subspecies. The<br />

distributional pattern of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl is more disjunct than that of <strong>the</strong> o<strong>the</strong>r<br />

subspecies, with <strong>the</strong> possible exception of <strong>the</strong><br />

Cali<strong>for</strong>nia spotted owl population in <strong>the</strong> mountain<br />

ranges of sou<strong>the</strong>rn Cali<strong>for</strong>nia (Noon and<br />

McKelvey 1992). The <strong>Mexican</strong> subspecies also<br />

appears to use a wider range of habitat types<br />

than <strong>the</strong> o<strong>the</strong>r subspecies. These unique aspects<br />

of <strong>the</strong> ecology of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

require unique approaches to its management.<br />

For example, threats to owl habitat and management<br />

proposed to address those threats may well<br />

differ among <strong>the</strong> diverse habitats occupied by<br />

<strong>Mexican</strong> spotted owls. In addition, because of<br />

its’ disjunct distributional pattern, dispersal<br />

among subpopulations of <strong>Mexican</strong> spotted owls<br />

is an important consideration. Thus, habitat<br />

management plans may need to consider not<br />

only areas occupied by owls but also intervening


areas, even where such areas are very different<br />

in habitat structure from those typically<br />

occupied by spotted owls.<br />

We have learned a great deal about <strong>the</strong><br />

<strong>Mexican</strong> spotted owl in <strong>the</strong> last decade, but<br />

significant in<strong>for</strong>mation gaps still remain. Most<br />

studies of <strong>the</strong> owl to date have been descriptive<br />

ra<strong>the</strong>r than experimental (III.D). Although<br />

we have identified patterns with<br />

Volume I/Part II<br />

35<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

respect to some aspects of <strong>the</strong> owls’ ecology (e.g.<br />

habitat use), cause and effect relationships have<br />

not been documented. Fur<strong>the</strong>r, many aspects of<br />

spotted owl demography and population structure<br />

remain unclear. These considerations<br />

suggest that much additional research is needed,<br />

and that management recommendations in <strong>the</strong><br />

near term must deal with high levels of uncertainty.


Volume I/Part II<br />

B. B. RECOVERY RECOVERY UNITS<br />

UNITS<br />

Sarah E. Rinkevich, Joseph L. Ganey, William H. Moir,<br />

Frank P. Howe, Fernando Clemente, and Juan F. Martinez-Montoya<br />

The <strong>Mexican</strong> spotted owl inhabits diverse<br />

<strong>for</strong>est types scattered across an even more physically<br />

diverse landscape. Fur<strong>the</strong>r, human activities<br />

vary dramatically throughout <strong>the</strong> owl’s range.<br />

These variations limit our ability to approach a<br />

status assessment on a rangewide basis. Consequently,<br />

we divided <strong>the</strong> range of <strong>the</strong> owl into<br />

11 geographic areas called “<strong>Recovery</strong> Units”<br />

(hereafter RUs). Six RUs were recognized within<br />

<strong>the</strong> United States: Colorado Plateau, Sou<strong>the</strong>rn<br />

Rocky Mountains - Colorado, Sou<strong>the</strong>rn Rocky<br />

Mountains - New Mexico, Upper Gila Mountains,<br />

Basin and Range - West, and Basin and<br />

Range - East (Figure. II.B.1). Five RUs were<br />

recognized in Mexico: Sierra Madre Occidental -<br />

Norte, Sierra Madre Oriental - Norte, Sierra<br />

Madre Occidental - Sur, Sierra Madre Oriental -<br />

Sur, and Eje Neovolcanico (Figure. II.B.2).<br />

UNITED UNITED STATES STATES<br />

STATES<br />

<strong>Recovery</strong> Units were identified based on <strong>the</strong><br />

following considerations (in order of importance):<br />

(1) physiographic provinces, (2) biotic<br />

regimes, (3) perceived threats to owls or <strong>the</strong>ir<br />

habitat, (4) administrative boundaries, and (5)<br />

known patterns of owl distribution. It is important<br />

to note that owl distributional patterns were<br />

a minor consideration in RU delineation, and<br />

that RUs do not necessarily represent discrete<br />

populations of owls. In fact, movement of<br />

individuals between RUs has been documented<br />

(Ganey and Dick 1995).<br />

Four major physiographic provinces were<br />

used in delineating RUs in <strong>the</strong> United States: <strong>the</strong><br />

Colorado Plateau, Basin and Range, Sou<strong>the</strong>rn<br />

Rocky Mountains, and Upper Gila Mountains<br />

(Hammond 1965, Wilson 1962, USGS 1970,<br />

Bailey 1980). Biotic regimes were based on<br />

classifications by Bailey (1980) and Brown et al.<br />

(1980). Administrative boundaries were used<br />

where management practices differed between<br />

jurisdictions (e.g., Sou<strong>the</strong>rn Rocky Mountains<br />

RUs). The following narratives describe dominant<br />

physical and biotic characteristics, patterns<br />

36<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

of owl distribution and habitat use, and <strong>the</strong><br />

dominant patterns of land ownership and land<br />

use within each RU.<br />

Colorado Colorado Plateau<br />

Plateau<br />

The Colorado Plateau RU (Figure. II.B.3)<br />

coincides with <strong>the</strong> Colorado Plateau Physiographic<br />

Province (USGS 1970). It includes<br />

most of south-central and sou<strong>the</strong>rn Utah plus<br />

portions of nor<strong>the</strong>rn Arizona, northwestern New<br />

Mexico, and southwestern Colorado. Major<br />

land<strong>for</strong>ms include interior basins and high<br />

plateaus dissected by deep canyons, including<br />

<strong>the</strong> canyons of <strong>the</strong> Colorado River and its<br />

tributaries (Williams 1986).<br />

Grasslands and shrub-steppes dominate <strong>the</strong><br />

Colorado Plateau at lower elevations, but woodlands<br />

and <strong>for</strong>ests dominate <strong>the</strong> higher elevations<br />

(Bailey 1980, West 1983). Pinyon pine and<br />

various juniper species comprise <strong>the</strong> primary tree<br />

types in <strong>the</strong> woodland zone. A montane zone<br />

extends over areas on <strong>the</strong> high plateaus and<br />

mountains (Bailey 1980). Forest types in this<br />

zone include ponderosa pine, mixed-conifer, and<br />

spruce-fir. Conifers may extend to lower elevations<br />

in canyons. Deciduous woody species<br />

dominate riparian communities, and are most<br />

common along major streams.<br />

The <strong>Mexican</strong> spotted owl reaches <strong>the</strong> northwestern<br />

limit of its range in this RU. <strong>Owl</strong><br />

habitat appears to be naturally fragmented in<br />

this RU, with most owls found in disjunct<br />

canyon systems or on isolated mountain ranges.<br />

In sou<strong>the</strong>rn Utah, breeding owls primarily<br />

inhabit deep, steep-walled canyons and hanging<br />

canyons. These canyons are typically surrounded<br />

by terrain that does not appear to support<br />

breeding spotted owls. <strong>Owl</strong>s also apparently<br />

prefer canyon terrain in southwestern Colorado,<br />

particularly in and around Mesa Verde National<br />

Park. In nor<strong>the</strong>rn Arizona and New Mexico,<br />

owls have been reported in both canyon and<br />

montane situations. Recent records of spotted<br />

owls exist <strong>for</strong> <strong>the</strong> Grand Canyon and Kaibab


Figur igur igure igur e II.B.1. II.B.1. <strong>Recovery</strong> Units within <strong>the</strong> United States.<br />

Volume I/Part II<br />

37<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Figur igur igure igur e II.B.2. II.B.2. II.B.2. <strong>Recovery</strong> Units within <strong>the</strong> Republic of Mexico.<br />

Volume I/Part II<br />

38<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Figur igur igure igur e II.B.3. II.B.3. Colorado Plateau <strong>Recovery</strong> Unit.<br />

Volume I/Part II<br />

39<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Plateau in Arizona, as well as <strong>for</strong> <strong>the</strong> Chuska<br />

Mountains, Black Mesa, Fort Defiance Plateau,<br />

and <strong>the</strong> Rainbow/Skeleton Plateau on <strong>the</strong><br />

Navajo Reservation. In addition, records exist <strong>for</strong><br />

<strong>the</strong> Zuni Mountains and Mount Taylor in New<br />

Mexico.<br />

Federal lands account <strong>for</strong> 44% of this RU<br />

(Table II.B.1). Tribal lands collectively total<br />

30%, with <strong>the</strong> largest single entity being <strong>the</strong><br />

Navajo Reservation. Private ownership accounts<br />

<strong>for</strong> 19%, and State lands just 8%. Most <strong>Mexican</strong><br />

spotted owls have been located on NPS lands in<br />

this RU, followed by FS and <strong>the</strong>n BLM lands<br />

(Ward et al. 1995).<br />

Recreation ranks first among land uses in<br />

National Parks within this RU. Activities such as<br />

hiking, camping, hunting, rock climbing, and<br />

mountain biking occur in owl habitat. Many of<br />

<strong>the</strong>se activities plus off-road vehicle recreation<br />

also occur on BLM and FS lands throughout <strong>the</strong><br />

Colorado Plateau. Various commercial enterprises<br />

relevant to industry and agriculture take<br />

place on <strong>the</strong>se lands. Particularly important are<br />

livestock grazing, timber cutting, coal and<br />

uranium mining, oil and natural gas pumping,<br />

and continued exploration <strong>for</strong> <strong>the</strong>se and o<strong>the</strong>r<br />

resources. Access roads, drill pads, pipelines, and<br />

loading and storage areas accompany all of <strong>the</strong>se<br />

activities.<br />

Sou<strong>the</strong>rn Sou<strong>the</strong>rn Rocky Rocky Mountains Mountains - - Colorado<br />

Colorado<br />

This RU (Figure. II.B.4) falls partly within<br />

<strong>the</strong> Sou<strong>the</strong>rn Rocky Mountains Physiographic<br />

Province (USGS 1970) and partly within <strong>the</strong><br />

Colorado Plateau Ecoregion (Bailey 1980). The<br />

Colorado - New Mexico state line delimits <strong>the</strong><br />

sou<strong>the</strong>rn boundary of this RU because land-use<br />

practices and potential threats on Federal lands<br />

differ between <strong>the</strong>se states. High mountain<br />

ranges characterize <strong>the</strong> RU (Curtis 1960);<br />

dominant ranges include <strong>the</strong> San Juan Mountains<br />

of southwestern Colorado, <strong>the</strong> Sangre de<br />

Cristo Mountains, and <strong>the</strong> Front Range.<br />

Vegetation ranges from grasslands at low<br />

elevations through pinyon-juniper woodlands,<br />

interior shrublands, ponderosa pine, mixedconifer<br />

and spruce-fir <strong>for</strong>ests, to alpine tundra<br />

on <strong>the</strong> highest peaks (Daubenmire 1943).<br />

Volume I/Part II<br />

40<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The <strong>Mexican</strong> spotted owl reaches <strong>the</strong> nor<strong>the</strong>astern<br />

limit of its range in this RU. Found<br />

primarily in canyons in this RU, <strong>the</strong> owls appear<br />

to occupy two disparate canyon habitat types.<br />

The first is sheer, slick-rock canyons containing<br />

widely scattered patches (up to 1 ha in size) of<br />

mature Douglas-fir in or near canyon bottoms or<br />

high on <strong>the</strong> canyon walls in short, hanging<br />

canyons. The second consists of steep canyons<br />

containing exposed bedrock cliffs ei<strong>the</strong>r close to<br />

<strong>the</strong> canyon floor or, more typically, several tiers<br />

of exposed rock at various heights on <strong>the</strong> canyon<br />

walls (Reynolds 1993). Mature Douglas-fir,<br />

white fir, and ponderosa pine dominate canyon<br />

bottoms and both north- and east-facing slopes.<br />

Ponderosa pine grows on <strong>the</strong> more xeric southand<br />

west-facing slopes, with pinyon-juniper<br />

growing on <strong>the</strong> mesa tops.<br />

Federal lands encompass 55% of <strong>the</strong> RU,<br />

with <strong>the</strong> majority administered by <strong>the</strong> FS,<br />

followed by <strong>the</strong> BLM and NPS (Table II.B.1).<br />

Approximately 40% of <strong>the</strong> land is privately<br />

owned, 3% is State administrated, and


41<br />

Table able II.B.1. II.B.1. Land ownership patterns (thousands of hectares) in <strong>Recovery</strong> Units (RU) within <strong>the</strong> United States.<br />

LAND AND ST STATUS ST TUS TUS CP CP1<br />

SRM-CO SRM-CO2<br />

SRM-NM SRM-NM3<br />

UGM UGM4<br />

BR-W BR-W5<br />

BR-E<br />

BR-E<br />

Federal Federal Lands<br />

Lands<br />

FS 2,503 5,546 1,220 3,520 2,238 495<br />

BLM 7,672 2,227 520 145 3,359 1,829<br />

NPS 1,678 140 14 17 172 59<br />

Total otal F FFederal<br />

F Federal<br />

ederal 11,853 11,853 11,853<br />

7,913 7,913<br />

1,754 1,754<br />

3,682 3,682<br />

5,769 5,769 2,383<br />

2,383<br />

State State Lands<br />

Lands<br />

AZ 957 0 0 20 2,905 0<br />

NM 300 0 200 192 84 856<br />

UT 862 0 0 0 0 0<br />

CO 11 454 0 0 0 0<br />

Total otal S SState<br />

S tate 2,129 2,129<br />

454 454<br />

200 200<br />

212 212<br />

2,989 2,989 856<br />

856<br />

Tribal Lands 8,026 85 463 941 1,922 382<br />

Private Lands 5,013 5,717 2,141 3,520 3,712 2,586<br />

O<strong>the</strong>r Lands 7 16 105 24 70 1,759 1,101<br />

TOTAL AL AL 27,037 27,037<br />

14,274 14,274<br />

4,582 4,582<br />

8,425 8,425<br />

16,151 16,151 7,308<br />

7,308<br />

1 Colorado Plateau RU<br />

2 Sou<strong>the</strong>rn Rocky Mountain - Colorado RU<br />

3 Sou<strong>the</strong>rn Rocky Mountain - New Mexico RU<br />

4 Upper Gila Mountains RU<br />

5 Basin and Range - West RU<br />

6 Basin and Range - East RU<br />

7 O<strong>the</strong>r Lands include U.S. Department of Defense, Bureau of Reclamation, U.S. Fish and Wildlife Service Refuge Lands, etc.<br />

BR-E 6


Figur igur igure igur e II.B.4. II.B.4. Sou<strong>the</strong>rn Rocky Mountains - Colorado <strong>Recovery</strong> Unit.<br />

Volume I/Part II<br />

42<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Figur igur igure igur e II.B.5. II.B.5. Sou<strong>the</strong>rn Rocky Mountains - New Mexico <strong>Recovery</strong> Unit.<br />

Volume I/Part II<br />

43<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Vegetation within <strong>the</strong> unit has been modified<br />

by past logging, grazing, surface mining,<br />

fuelwood ga<strong>the</strong>ring, and fire suppression (Williams<br />

1986, Van Hooser et al. 1993). Ponderosa<br />

pine, mixed-conifer, and spruce-fir <strong>for</strong>ests are<br />

widespread at higher elevations. Juniper savanna<br />

and montane grasslands dominate lower elevations<br />

(Brown et al. 1980). In some areas, mesa<br />

tops dominated by ponderosa pine and juniper<br />

are dissected by steep canyons. Vegetation on<br />

canyon slopes and bottoms includes a variety of<br />

coniferous and deciduous trees.<br />

In general, owls inhabit steep terrain and<br />

canyons in this RU. They typically occur in<br />

mixed-conifer <strong>for</strong>ests on steep slopes in <strong>the</strong><br />

Sangre de Cristo Mountains, and in <strong>the</strong> Jemez<br />

Mountains <strong>the</strong>y occupy canyons incised into<br />

volcanic rock. Patches of mixed-conifer <strong>for</strong>est<br />

which appear to contain attributes of owl habitat<br />

exist throughout nor<strong>the</strong>rn New Mexico.<br />

Privately owned lands comprise 47% of <strong>the</strong><br />

total land within this RU (Table II.B.1). Federal<br />

lands account <strong>for</strong> 38%, numerous Pueblos and<br />

Tribal lands 10%, and State-administered lands<br />

4%. <strong>Mexican</strong> spotted owls have been found<br />

primarily on FS lands, with several records in<br />

Bandelier National Monument as well (Johnson<br />

and Johnson 1985:5).<br />

Dominant land-use practices within this RU<br />

include timber cutting and livestock grazing.<br />

Products such as vigas (small- to mediumdiameter<br />

trees, generally 30-35cm dbh, used <strong>for</strong><br />

traditional southwest ceiling beams), latillas<br />

(small-diameter trees, generally 10cm dbh apsen<br />

saplings, used <strong>for</strong> decorative southwest ceilings<br />

or fences), and fuelwood are harvested <strong>for</strong><br />

personal use. Recreational activities in nor<strong>the</strong>rn<br />

New Mexico include skiing, off-road driving,<br />

hiking, camping, and hunting. O<strong>the</strong>r land uses<br />

include oil, natural gas, and mineral development,<br />

and pipeline corridors.<br />

Upper Upper Gila Gila Mountains<br />

Mountains<br />

The Upper Gila Mountains RU (Figure.<br />

II.B.6) is based on <strong>the</strong> Upper Gila Mountains<br />

Forest Province (Bailey 1980). Williams (1986)<br />

refers to this area as <strong>the</strong> Datil-Mogollon Section,<br />

part of a physiographic subdivision transitional<br />

Volume I/Part II<br />

44<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

between <strong>the</strong> Basin and Range and Colorado<br />

Plateau Provinces. This complex area consists<br />

of steep mountains and deep entrenched river<br />

drainages dissecting high plateaus. The<br />

Mogollon Rim, a prominent fault scarp, bisects<br />

<strong>the</strong> unit.<br />

McLaughlin (1986) described a “Mogollon”<br />

floral element in this region. The vegetation is a<br />

zonal pattern of grasslands at lower elevations<br />

upward through pinyon-juniper woodlands,<br />

ponderosa pine, mixed-conifer, and spruce-fir<br />

<strong>for</strong>ests at higher elevations. Many canyons<br />

contain stringers of deciduous riparian <strong>for</strong>ests,<br />

particularly at low and middle elevations. This<br />

unit contains <strong>the</strong> largest contiguous ponderosa<br />

pine <strong>for</strong>est in North American, an unbroken<br />

band of <strong>for</strong>est 25 to 40 miles wide and approximately<br />

300 miles long extending from northcentral<br />

Arizona to west-central New Mexico<br />

(Cooper 1960).<br />

<strong>Mexican</strong> spotted owls are widely distributed<br />

and use a variety of habitats within <strong>the</strong> Upper<br />

Gila Mountains RU. <strong>Owl</strong>s are most common in<br />

mixed-conifer <strong>for</strong>ests dominated by Douglas-fir<br />

and/or white fir and canyons with varying<br />

degrees of <strong>for</strong>est cover (Ganey and Balda 1989a,<br />

Ganey and Dick 1995). <strong>Owl</strong>s also occur in<br />

ponderosa pine-Gambel oak <strong>for</strong>est, where <strong>the</strong>y<br />

are typically found in stands containing welldeveloped<br />

understories of Gambel oak.<br />

Federal lands, mostly FS, encompass 44% of<br />

this RU (Table II.B.1). Tribal lands account <strong>for</strong><br />

11%, privately owned lands 42%, and State<br />

lands 3%. The greatest concentration of <strong>the</strong><br />

known <strong>Mexican</strong> spotted owl population occurs<br />

within this RU, and most known owl locations<br />

occur on FS and Tribal lands (Ward et al. 1995).<br />

Many spotted owls are found within wilderness<br />

areas in this RU with <strong>the</strong> Gila Wilderness<br />

supporting <strong>the</strong> largest known wilderness population.<br />

The major land use within this RU is timber<br />

harvest. All of <strong>the</strong> National Forests as well as <strong>the</strong><br />

Fort Apache and San Carlos Indian Reservations<br />

have active timber management programs.<br />

Fuelwood harvest, including both personal and<br />

commercial harvest, occurs across much of this<br />

unit. Livestock grazing is ubiquitous on FS lands<br />

and widespread over large portions of <strong>the</strong> Fort<br />

Apache and San Carlos Indian Reservations. In


45<br />

Figur igur igure igur e II.B.6. II.B.6. Upper Gila Mountains <strong>Recovery</strong> Unit.


addition, recreational activities such as hiking,<br />

camping, and hunting attract many people to<br />

this RU.<br />

Basin Basin Basin and and Range Range - - West West<br />

West<br />

The Basin and Range Area Province (USGS<br />

1970, Bailey 1980) provided <strong>the</strong> basis <strong>for</strong> two<br />

RUs (Figure. II.B.7). We subdivided <strong>the</strong> Basin<br />

and Range area into eastern and western units<br />

using <strong>the</strong> Continental Divide as <strong>the</strong> partition<br />

between <strong>the</strong>se units. The division was based on<br />

differences in climatic and floristic characteristics<br />

between <strong>the</strong>se areas. The Basin and Range - West<br />

flora is dominated by Madrean elements while<br />

<strong>the</strong> Basin and Range - East unit shows more<br />

Rocky Mountain affinities (Brown et al. 1980).<br />

Geologically, <strong>the</strong> Basin and Range - West<br />

RU exhibits horst and graben faulting (Wilson<br />

1962) with numerous fault-block mountains<br />

separated by valleys. Complex faulting and<br />

canyon carving define <strong>the</strong> physical landscape<br />

within <strong>the</strong>se mountains. These ranges include,<br />

but are not limited to, <strong>the</strong> Chiricahua,<br />

Huachuca, Pinaleno, Bradshaw, Pinal, Santa<br />

Catalina, Santa Rita, Patagonia, Santa Teresa,<br />

Atascosa, Mule, Dragoon, Peloncillo, Mazatzal,<br />

and Rincon Mountains.<br />

Vegetation ranges from desert scrubland and<br />

semi-desert grassland in <strong>the</strong> valleys upwards to<br />

montane <strong>for</strong>ests. Montane vegetation includes<br />

interior chaparral, encinal woodlands, and<br />

Madrean pine-oak woodlands at low and middle<br />

elevations, with ponderosa pine, mixed-conifer,<br />

and spruce-fir <strong>for</strong>ests at higher elevations (Brown<br />

et al. 1980). Isolated mountain ranges are<br />

surrounded by Sonoran and Chihuahuan desert<br />

basins.<br />

<strong>Mexican</strong> spotted owls occupy a wide range of<br />

habitat types within this RU. The majority of<br />

owls occur in isolated mountain ranges where<br />

<strong>the</strong>y inhabit encinal oak woodlands, mixedconifer<br />

and pine-oak <strong>for</strong>ests, and rocky canyons<br />

(Ganey and Balda 1989a, Duncan and Taiz<br />

1992, Ganey et al. 1992).<br />

Federal lands encompass 36% of this RU,<br />

mostly administered by <strong>the</strong> BLM followed by<br />

<strong>the</strong> FS and a small portion by <strong>the</strong> NPS (Table<br />

II.B.1). Privately owned lands amount to 22%,<br />

State lands 19%, Tribal lands (San Carlos<br />

Volume I/Part II<br />

46<br />

Apache Reservation) 12%, and DOD lands<br />

11%. Within this RU <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl occupies primarily FS lands, and <strong>the</strong> majority<br />

occur within <strong>the</strong> Coronado National Forest.<br />

DOD lands also support <strong>the</strong> owl on Fort<br />

Huachuca Army Base in <strong>the</strong> Huachuca<br />

Mountains.<br />

Recreation dominates land use within this<br />

unit. Activities such as hiking, birdwatching,<br />

camping, off-road driving, skiing, and hunting<br />

are particularly popular. Livestock grazing is<br />

widespread but most intensive at low and middle<br />

elevations. Urban and rural development and<br />

mining modify portions of <strong>the</strong> Basin and Range<br />

- West landscape. Timber harvest occurs mainly<br />

on <strong>the</strong> Prescott National Forest and <strong>the</strong> San<br />

Carlos Apache Indian Reservation. According to<br />

<strong>the</strong> Coronado National Forest Land Management<br />

<strong>Plan</strong>, timber cutting is used sparingly to<br />

enhance wildlife and recreational values. Military<br />

training maneuvers take place in and around<br />

<strong>Mexican</strong> spotted owl habitat on Fort Huachuca<br />

Army Base.<br />

Basin Basin and and and Range Range - - East<br />

East<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

We delineated <strong>the</strong> Basin and Range - East<br />

RU (Figure. II.B.8) based on <strong>the</strong> Basin and<br />

Range Area Province (USGS 1970) and <strong>the</strong><br />

Desert and Steppic Ecoregions (Bailey 1980).<br />

This RU is characterized by numerous parallel<br />

mountain ranges separated by alluvial valleys and<br />

broad, flat basins. Williams (1986) refers to <strong>the</strong><br />

Rio Grande Rift as <strong>the</strong> separation between <strong>the</strong><br />

Basin and Range physiographic province and <strong>the</strong><br />

Colorado Plateau and Upper Gila Mountains<br />

physiographic provinces. The climate features<br />

mild winters, as indicated by <strong>the</strong> presence of<br />

broad-leaved evergreen plants at relatively high<br />

elevations (USDA 1991).<br />

Regional vegetation ranges from Chihuahuan<br />

desert scrubland and Great Basin grasslands<br />

at low elevations, through Great Basin<br />

woodland (pinyon-juniper) at middle elevations<br />

to petran montane coniferous <strong>for</strong>ests at high<br />

elevations (Brown et al. 1980). Montane habitat<br />

includes ponderosa pine, mixed-conifer, and<br />

spruce-fir <strong>for</strong>ests and is patchily distributed<br />

throughout <strong>the</strong> higher mountain ranges. Cottonwood<br />

bosques as well as o<strong>the</strong>r riparian


Figur igur igure igur e II.B.7. II.B.7. Basin and Range - West <strong>Recovery</strong> Unit.<br />

Volume I/Part II<br />

47<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Figur igur igure igur e II.B.8. II.B.8. II.B.8. Basin and Range - East <strong>Recovery</strong> Unit.<br />

Volume I/Part II<br />

48<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


vegetation exist along <strong>the</strong> Rio Grande corridor.<br />

Montane and especially riparian communities<br />

have been altered considerably by human activities.<br />

<strong>Mexican</strong> spotted owls occur in <strong>the</strong> isolated<br />

mountain ranges scattered across this RU. They<br />

are most common in mixed-conifer <strong>for</strong>est but<br />

are also found in ponderosa pine <strong>for</strong>est and<br />

pinyon-juniper woodland (Skaggs and Raitt<br />

1988). The owl has been found within mixedconifer<br />

canyon habitat in <strong>the</strong> Guadalupe Mountains<br />

(McDonald et al. 1991).<br />

Of <strong>the</strong> Basin and Range - East RU land area,<br />

private lands encompass 35%, Federal lands<br />

48%, State lands 12%, and Tribal lands 5%<br />

(Table II.B.1). The Mescalero Apache Indian<br />

Reservation comprises <strong>the</strong> largest portion of <strong>the</strong><br />

Tribal lands. The majority of known <strong>Mexican</strong><br />

spotted owls are located on FS lands, with some<br />

found on NPS and Tribal lands.<br />

Dominant land uses within this RU include<br />

timber management and livestock grazing.<br />

Recreational activities such as off-road driving,<br />

skiing, hiking, camping, and hunting are also<br />

locally common within <strong>the</strong> RU.<br />

Volume I/Part II<br />

MEXICO<br />

MEXICO<br />

Conserving its natural resources has been a<br />

significant challenge <strong>for</strong> Mexico. To meet <strong>the</strong><br />

challenge, <strong>the</strong> National System of Protected<br />

Areas was <strong>for</strong>med; in March of 1988, <strong>the</strong> General<br />

Law of Ecological Balance and Environmental<br />

Protection was implemented. A total of 5,992<br />

km 2 (almost 600,000 ha) has been decreed as<br />

Protected Natural Areas within <strong>the</strong> RUs. This<br />

expanse has been classified into nine categories<br />

according to <strong>the</strong> management objectives and <strong>the</strong><br />

legal uses of particular areas. The categories<br />

include: (1) Biosphere Reserves, (2) Special<br />

Biosphere Reserves, (3) National Parks, (4)<br />

National Monuments, (5) National Marine<br />

Parks, (6) Areas of Protection of Natural Resources,<br />

(7) Areas of Protection of Land and<br />

Aquatic Wildlife, (8) Urban Parks, and (9) Areas<br />

Subject to Ecological Conservation. Overall,<br />

<strong>the</strong>re are three types of land tenancy exist in<br />

Mexico: (1) Federal lands, which include<br />

different institutions of <strong>the</strong> Federal Government<br />

such as Protected Natural Areas; (2) ejidal land,<br />

49<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

which includes land allotted by <strong>the</strong> <strong>Mexican</strong><br />

Government to a person or community, <strong>for</strong><br />

agriculture, <strong>for</strong>estry, mining, or o<strong>the</strong>r uses; and<br />

(3) private land.<br />

The five RUs in Mexico include Sierra<br />

Madre Occidental - Norte, Sierra Madre Oriental<br />

- Norte, Sierra Madre Occidental -Sur, Sierra<br />

Madre Oriental - Sur, and Eje Neovolcanico<br />

(Figure. II.B.2). Three major physiographic<br />

provinces were used in <strong>the</strong> delineation: Sierra<br />

Madre Occidental, Sierra Madre Oriental, and<br />

Sistema Volcanico Transversal (Cuanalo et al.<br />

1989). Criteria used to delineate RUs in Mexico<br />

were similar to that used to con<strong>for</strong>m <strong>the</strong> RUs in<br />

<strong>the</strong> United States. These criteria, listed in order<br />

of importance, were: (1) distribution of <strong>the</strong><br />

spotted owl, (2) local vegetation, (3) physiographic<br />

features, (4) administrative boundaries,<br />

and (5) potential threats to <strong>the</strong> conservation of<br />

<strong>the</strong> owl and its habitat.<br />

<strong>Owl</strong> distribution is disjunct across Mexico.<br />

Williams and Skaggs (1993) report spotted owls<br />

at 53 locations in 11 mainland <strong>Mexican</strong> States.<br />

Although vegetation types differ throughout<br />

each RU, oak and pine-oak <strong>for</strong>est types appeared<br />

to be commonly associated with owl habitat in<br />

most or all RUs. These oak species included<br />

Quercus resinosa, Q. gentryi, Q. eduardii, Q.<br />

grisea, Q. chihuahuensis, Q. potosina/Q. laeta, and<br />

Q. coccolobifolia. Fur<strong>the</strong>r, Pinus teocote was <strong>the</strong><br />

most common pine occurring on upper mesas<br />

and occasionally on north-facing slopes in some<br />

areas where owls were found. Land uses within<br />

all RUs include timber cutting, cattle and sheep<br />

grazing, fuelwood ga<strong>the</strong>ring, and clearing <strong>for</strong>ested<br />

areas <strong>for</strong> agriculture. Although, <strong>the</strong>se land<br />

uses are practiced in different amounts throughout<br />

each RU, <strong>the</strong> majority occur within ejidos.<br />

The following narratives describe dominant<br />

physical and biotic attributes, distribution of<br />

owls, and land administration and ownership<br />

of each unit.<br />

Sierra Sierra Madre Madre Occidental Occidental - - Norte<br />

Norte<br />

Covering an enormous area, <strong>the</strong> Sierra<br />

Madre Occidental - Norte includes parts of <strong>the</strong><br />

States of Chihuahua, Sinaloa, Durango, and<br />

Sonora. In general, this area is characterized by<br />

isolated mountain ranges surrounded by both


narrow and wide valleys. Vegetation communities<br />

consist of pine-oak <strong>for</strong>est, tropical deciduous<br />

<strong>for</strong>est, oak <strong>for</strong>est, microphyll shrub, and grassland.<br />

<strong>Mexican</strong> spotted owls have been reported in<br />

<strong>the</strong> nor<strong>the</strong>rn and western portions of this RU. A<br />

recent study in Sonora found 12 sites in isolated<br />

mountain ranges (Cirett-Galan and Diaz 1993).<br />

The owls occupied canyons and slopes with<br />

various exposures, and most were found in pineoak<br />

<strong>for</strong>est. In portions of Chihuahua, 25 owls<br />

were located at 13 different localities in several<br />

mountain ranges (Tarango et al. 1994). Most<br />

owls were found in small, isolated patches of<br />

pine-oak <strong>for</strong>est in canyons.<br />

Records <strong>for</strong> <strong>the</strong> State of Sinaloa are limited.<br />

There are at least two records from <strong>the</strong> high<br />

Rancho Liebre Barranca, near <strong>the</strong> Sinaloa-<br />

Durango State line (Williams and Skaggs 1993).<br />

These sites were described as deep canyons<br />

containing pine-oak and subtropical vegetation<br />

(Alden 1969).<br />

Private lands comprise 74%, ejidos 25%, and<br />

Federal lands 1% of <strong>the</strong> total land within this<br />

RU (Table II.B.2). Chihuahua has two National<br />

Parks: Cascadas de Bassaseachic, and Cumbres<br />

de Majalca. This RU also includes La Michilia<br />

Biosphere Reserve, located in Durango. Biosphere<br />

Reserves are protected areas with<br />

relatively unaltered landscapes and contain<br />

endemic, threatened, or endangered species.<br />

Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Sierra Sierra Madre Madre Oriental Oriental - - Norte Norte<br />

Norte<br />

Table able II.B.2. II.B.2. Land ownership patterns (thousands of hectares) in <strong>Recovery</strong> Units within Mexico.<br />

__________________________________________________________________________________________<br />

__________________________________________________________________________________________<br />

Land Land Land Ownership Ownership Ownership SMO SMO SMOcN SMO cN 1 SMO SMOrN SMO rN 2 SMO SMOcS SMO SMOcS<br />

cS 3 SMO SMO SMOrS SMO SMOrS<br />

rS 4 ENV ENV5<br />

_________________________________________________________________________________________<br />

_________________________________________________________________________________________<br />

Ejidos6 4,783 28 1,220 235 441<br />

PNAs7 46 0 38 250 274<br />

Private 14,100 7,506 2,075 1,630 5,306<br />

Total otal 18,929 18,929 7,534 7,534 3,333 3,333 2,115 2,115<br />

6,021<br />

6,021<br />

1 Sierra Madre Occidental - Norte RU<br />

2 Sierra Madre Oriental - Norte RU<br />

3 Sierra Madre Occidental - Sur RU<br />

4 Sierra Madre Oriental - Sur RU<br />

5 Eje Neovolcanico RU<br />

6 Ejidos are lands allocated by <strong>the</strong> <strong>Mexican</strong> Government to a person or community, to be<br />

used <strong>for</strong> agriculture, <strong>for</strong>estry, mining, etc.<br />

7 Protected Natural Areas<br />

50<br />

The Sierra Madre Oriental - Norte includes<br />

<strong>the</strong> central portion of <strong>the</strong> State of Coahuila. This<br />

area is characterized by broad mountain ranges<br />

surrounded by valleys. Vegetation consists of<br />

grasslands, mesquite woodland, dwarf oak<br />

groves, submontane shrubland, desert shrubland,<br />

crasicaule shrub, and pine-oak and oak <strong>for</strong>ests.<br />

Two owl records are reported <strong>for</strong> this RU. At<br />

one of <strong>the</strong>se sites an owl was observed roosting<br />

in a canyon bottom under a dense canopy of<br />

maples and oaks. Vegetation in <strong>the</strong> o<strong>the</strong>r canyon<br />

was described as “garden-like,” containing pines,<br />

oaks, and madrones (Williams and Skaggs<br />

1993).<br />

Lands in this RU are almost entirely privately<br />

owned (Table II.B.2). Private lands encompass<br />

over 99% and ejidos comprise < 1% of<br />

<strong>the</strong> total land area. No Protected Natural Areas<br />

of any category exist in this RU.<br />

Sierra Sierra Madre Madre Occidental Occidental Occidental - - Sur<br />

Sur<br />

The Sierra Madre Occidental - Sur RU<br />

includes parts of <strong>the</strong> States of Durango,<br />

Zacatecas, San Luis Potosi, Aguascalientes,<br />

Jalisco, Nayarit, Queretaro, and Guanajuato. In<br />

general, this area is characterized by isolated


mountains, valleys, and severely dissected canyons<br />

and gorges. Vegetation includes mesquite<br />

woodland, submontane shrub, grasslands, pineoak<br />

<strong>for</strong>est, crasicaule shrub, low tropical deciduous<br />

<strong>for</strong>est, and desert shrubland.<br />

Records exist <strong>for</strong> owls in La Michilia Biosphere<br />

Reserve. In addition, <strong>Mexican</strong> spotted<br />

owls have recently been found in Aguascalientes<br />

near <strong>the</strong> border of Zacatecas, in <strong>the</strong> Sierra Fria<br />

(Williams and Skaggs 1993). <strong>Owl</strong> records also<br />

exist within Guanajuato State.<br />

Private lands comprise 62%, ejidos 37%, and<br />

Federal lands 1% of this RU (Table II.B.2).<br />

Federal lands include two National Parks: El<br />

Climatario in Queretaro State, and Gogorron in<br />

<strong>the</strong> State of San Luis Potosi. In addition, this RU<br />

includes Mariposa Monarca Sanctuary, a Special<br />

Biosphere Reserve. The Special Biosphere<br />

Reserves have one or more ecosystems, are<br />

relatively unaltered by anthropogenic activities,<br />

and contain endemic, threatened, or endangered<br />

species.<br />

Sierra Sierra Madre Madre Oriental Oriental Oriental - - Sur<br />

Sur<br />

The Sierra Madre Oriental - Sur includes<br />

parts of <strong>the</strong> States of Coahuila, Nuevo Leon, and<br />

Tamaulipas. This RU is characterized by long<br />

ridges with sharp pinnacles, narrow valleys, and a<br />

few plateaus. Vegetation consists of pine <strong>for</strong>est,<br />

submontane shrublands, dwarf oak, and desert<br />

rosetofilo shrublands.<br />

<strong>Mexican</strong> spotted owls have been found in<br />

<strong>the</strong> sou<strong>the</strong>rn portions of Coahuila (Williams and<br />

Skaggs 1993) and in Tamaulipas (Ward et al.<br />

1995). The owls were found in oak, pine,<br />

juniper, and mixed-conifer <strong>for</strong>ests. They were<br />

reported to use cliff sites <strong>for</strong> nesting and roosting.<br />

Five locations have been reported in Nuevo<br />

Leon. These sites were described as pine-oak and<br />

mixed-conifer <strong>for</strong>ests with large cliffs having<br />

nor<strong>the</strong>ast exposures.<br />

Volume I/Part II<br />

51<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

This RU is comprised of 77% private property,<br />

12% Federal lands, and 11% ejidos (Table<br />

II.B.2). The Federal lands include one National<br />

Park, Cumbres de Monterrey in Nuevo Leon.<br />

Natural Monument Cerro de la Silla, in Nuevo<br />

Leon, is also within this RU. Natural Monuments<br />

possess one or more elements of national<br />

significance. These elements may be sites or<br />

natural objects that have been placed under<br />

absolute protection because of <strong>the</strong>ir unique and<br />

exceptional makeup, aes<strong>the</strong>tic interest, and/or<br />

historical or scientific value.<br />

Eje Eje Neovolvanico<br />

Neovolvanico<br />

The Eje Neovolcanico RU covers portions<br />

of many States including Jalisco, Michoacan,<br />

Guanajuato, Queretaro, Hidalgo, Mexico,<br />

Guerrero, Puebla, Morelaos, Tlaxcala Veracruz,<br />

and Oaxaca. This RU is characterized by volcanic<br />

cones severely dissected by ravines. The<br />

area also includes rounded hills, slopes, and<br />

plateaus. Vegetation communities include pineoak<br />

<strong>for</strong>est, grassland, low tropical deciduous<br />

<strong>for</strong>est, crasicaule shrub, oak <strong>for</strong>est, juniper <strong>for</strong>est,<br />

pine <strong>for</strong>est, mesquite woodlands, and desert<br />

shrublands.<br />

<strong>Mexican</strong> spotted owls have been reported in<br />

Jalisco on <strong>the</strong> volcano of Cerro Nevado de<br />

Colima (Voacan de Nieve). Vegetation in this<br />

area consists of pine-oak <strong>for</strong>est. One <strong>Mexican</strong><br />

spotted owl was collected near <strong>the</strong> city of<br />

Uruapan in <strong>the</strong> State of Michoacan at Cerro de<br />

Tancitaro. However, this area is now urbanized<br />

and no longer contains owl habitat. Although<br />

o<strong>the</strong>r states in this RU appear to contain suitable<br />

owl habitat, Jalisco is <strong>the</strong> only State known to<br />

have recent records of spotted owls.<br />

This unit is comprised of 88% private lands,<br />

5% Federal lands, and 7% ejidos (Table II.B.2).<br />

This RU includes 19 National Parks, 1 Special<br />

Biosphere Reserve, and 1 Area of Protection of<br />

Land and Aquatic Wildlife.


Volume I/Part II<br />

C. C. DEFINITIONS DEFINITIONS OF OF FOREST FOREST COVER COVER TYPES<br />

TYPES<br />

James L. Dick, Jr., Joseph L. Ganey, and William H. Moir<br />

This <strong>Recovery</strong> <strong>Plan</strong> proposes specific<br />

guidelines <strong>for</strong> several <strong>for</strong>est cover types, including<br />

mixed-conifer, pine-oak, and riparian <strong>for</strong>ests<br />

(III.B). This is based on: (1) considerable evidence<br />

that <strong>the</strong>se cover types are of specific<br />

importance to <strong>the</strong> <strong>Mexican</strong> spotted owl in terms<br />

of providing habitat <strong>for</strong> nesting, roosting, and<br />

<strong>for</strong>aging activities (Ganey and Dick 1995); and<br />

(2) <strong>the</strong> Team’s desire to target guidelines <strong>for</strong> <strong>the</strong><br />

most appropriate habitats and avoid imposing<br />

restrictions where specific guidelines to protect<br />

spotted owl habitat are unwarranted.<br />

Numerous treatments deal with <strong>the</strong> concepts<br />

of classifying vegetation to cover or habitat types<br />

(e.g., Daubenmire 1952, 1968, Pfister 1989).<br />

These concepts will not be reviewed in any<br />

depth here. In general, we accept <strong>the</strong> view that<br />

<strong>the</strong> basic unit of classification of climax vegetation<br />

is <strong>the</strong> plant association (Küchler 1964,<br />

Daubenmire 1968, Pfister 1989). These associations<br />

are defined using in<strong>for</strong>mation on present<br />

species composition and successional pathways.<br />

The problem with applying guidelines to plant<br />

associations is that many <strong>for</strong>ests in <strong>the</strong> southwest<br />

may and should not be in or even near a climax<br />

condition because of <strong>the</strong> frequency and intensity<br />

of disturbance events in <strong>the</strong>se <strong>for</strong>ests. For example,<br />

in an analysis of <strong>Mexican</strong> spotted owl<br />

habitat on <strong>the</strong> Alpine Ranger District, Apache-<br />

Sitgreaves National Forest, <strong>the</strong> Team determined<br />

that habitat classifications based on current and<br />

climax vegetation gave very different results.<br />

Based on current vegetation, important roosting<br />

and nesting habitat typed out as mixed-conifer<br />

<strong>for</strong>est, whereas a classification based on potential<br />

natural vegetation (PNV) typed many of <strong>the</strong>se<br />

areas as spruce-fir <strong>for</strong>est. This points out <strong>the</strong><br />

need <strong>for</strong> clear, operational definitions of cover<br />

types to be used when applying guidelines under<br />

this <strong>Plan</strong>.<br />

In this section, we first review some of <strong>the</strong><br />

relevant literature on <strong>for</strong>est cover types in <strong>the</strong><br />

southwest, and <strong>the</strong>n provide operational definitions<br />

and a simple key that should allow land<br />

managers to classify lands in a manner compatible<br />

with <strong>the</strong> recommendations provided in this<br />

52<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>Plan</strong>. Our intent is not to provide a comprehensive<br />

classification scheme here or to supplant<br />

extant classification schemes. Ra<strong>the</strong>r, our intent<br />

is to provide guidance to land managers charged<br />

with applying <strong>Recovery</strong> <strong>Plan</strong> guidelines and to<br />

facilitate uni<strong>for</strong>m application of guidelines across<br />

administrative boundaries.<br />

LITERATURE LITERATURE REVIEW<br />

REVIEW<br />

Extensive literature exists on both vegetation<br />

classification in general and on classification<br />

systems <strong>for</strong> southwestern <strong>for</strong>ests. Our intent is<br />

not to review that literature exhaustively, but to<br />

present an overview of some classification<br />

systems currently in use. This in<strong>for</strong>mation will<br />

provide <strong>the</strong> background <strong>for</strong> discussion of <strong>for</strong>est<br />

type definitions relevant to this <strong>Plan</strong>.<br />

Soil temperature (STR) and moisture (SMR)<br />

regimes provide one possible approach to classifying<br />

<strong>for</strong>est types. STR and SMR may be used to<br />

conceptualize three major groups of cover types<br />

in <strong>the</strong> southwestern United States. Ponderosa<br />

pine <strong>for</strong>ests typically occur where <strong>the</strong> STR is<br />

frigid or (in sou<strong>the</strong>rn Arizona and southwestern<br />

New Mexico) mesic and where <strong>the</strong> SMR is ustic.<br />

Spruce-fir <strong>for</strong>ests everywhere occur on soils of<br />

cryic SMR, whereas mixed-conifer <strong>for</strong>ests are<br />

uniquely udic and frigid in <strong>the</strong>ir SMR and STR,<br />

respectively. The implication is that <strong>the</strong>se soil<br />

parameters partition <strong>the</strong> soil environment into<br />

three mutually exclusive but all encompassing<br />

classes (USDA 1991). These classes are generally<br />

consistent with <strong>the</strong> three major <strong>for</strong>est type<br />

groups mentioned.<br />

Most vegetation-classification schemes,<br />

however, are based on ei<strong>the</strong>r exisiting vegetation<br />

or on a combination of existing vegetation and<br />

knowledge of successional potential (Layser and<br />

Schubert 1979). Eyre (1980) discussed <strong>the</strong><br />

practice of defining <strong>for</strong>est cover types on <strong>the</strong><br />

basis of “present occupancy of an area by tree<br />

species.” He fur<strong>the</strong>r described <strong>the</strong> practice of<br />

naming <strong>for</strong>est types after <strong>the</strong> dominant tree<br />

species. Dominance was determined by relative<br />

proportions of basal area, and <strong>the</strong> type name was


usually confined to one or two species. An added<br />

requirement was that a species must contribute<br />

at least 20% of <strong>the</strong> total basal area to be used in<br />

<strong>the</strong> type name.<br />

Numerous authors have expanded on this<br />

approach, and developed and refined classification<br />

systems <strong>for</strong> southwestern <strong>for</strong>ests based not<br />

only on existing vegetation but also on estimated<br />

site potential. These treatments are discussed<br />

below.<br />

Ponderosa Ponderosa Pine<br />

Pine<br />

The ponderosa pine <strong>for</strong>est type occurs in<br />

what Moir (1993) described as <strong>the</strong> Lower Montane<br />

Coniferous Forest. Forests in this zone are<br />

dominated by pines, sometimes co-occurring<br />

with junipers and oaks. The climate is sometimes<br />

borderline <strong>for</strong> <strong>for</strong>ests, with moisture becoming<br />

limiting in <strong>the</strong> upper portions of <strong>the</strong> soil profile<br />

during part of <strong>the</strong> long growing season. Moir<br />

(1993) included <strong>the</strong> following series in this<br />

general <strong>for</strong>est type: Ponderosa pine - Gambel<br />

oak, Ponderosa pine - silverleaf oak, Ponderosa<br />

pine - pinyon pine - Gambel oak, Ponderosa<br />

pine - pinyon pine - gray oak, and Chihuahua<br />

pine.<br />

Layser and Schubert (1979) described <strong>the</strong><br />

Ponderosa Pine Series as being generally dominated<br />

by <strong>the</strong> Rocky Mountain variety of ponderosa<br />

pine (var. scopulorum), except in sou<strong>the</strong>astern<br />

Arizona, where Pinus ponderosa var. arizonica<br />

dominates. This series occurs in areas that are<br />

generally too warm or too dry <strong>for</strong> Douglas-fir<br />

and/or true firs. Gambel oak is often a long-lived<br />

seral species. The Ponderosa Pine Series in <strong>the</strong><br />

southwest is generally more complex than that<br />

described <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn Rocky Mountains,<br />

because of <strong>the</strong> additional associated tree species<br />

and <strong>the</strong> presence of two varieties of ponderosa<br />

pine (Layser and Schubert 1979). Hanks et al.<br />

(1983), Alexander et al. (1984a), Alexander and<br />

Ronco (1987), DeVelice et al. (1986), and<br />

Fitzhugh et al. (1987) provide fur<strong>the</strong>r discussion<br />

of <strong>the</strong> Ponderosa Pine Series and associated<br />

habitat types and phases in <strong>the</strong> southwestern<br />

United States.<br />

Volume I/Part II<br />

53<br />

Mixed-Conifer<br />

Mixed-Conifer<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Mixed-conifer <strong>for</strong>ests in <strong>the</strong> southwestern<br />

United States generally approximate to <strong>the</strong><br />

Upper Montane Coniferous Forest discussed by<br />

Moir (1993). Mixed-conifer <strong>for</strong>ests are most<br />

common between approximately 2,440 and<br />

3,050 m (8,000-10,000 feet) in elevation, but<br />

may occur higher or lower depending on topography<br />

and aspect. In particular, mixed-conifer<br />

<strong>for</strong>est may extend to lower elevations in canyon<br />

systems and cold-air drainages.<br />

Southwestern mixed-conifer <strong>for</strong>ests are<br />

among <strong>the</strong> most complex <strong>for</strong>est types known,<br />

exhibiting great variation in tree composition<br />

(USDA 1983). Overstory species in <strong>the</strong>se <strong>for</strong>ests<br />

include Rocky Mountain Douglas-fir, white fir,<br />

Rocky Mountain ponderosa pine, quaking<br />

aspen, southwestern white pine, limber pine,<br />

and blue spruce. Forests in any successional stage<br />

may be mixed-conifer if tree regeneration indicates<br />

any of <strong>the</strong> above tree species will assume<br />

dominance in time. Some stands may consist of<br />

only two species, whereas o<strong>the</strong>rs may contain as<br />

many as eight associates (USDA 1983). Gambel<br />

oak and/or silverleaf oak may share overstory or<br />

understory dominance with <strong>the</strong> conifers in<br />

mixed-conifer <strong>for</strong>ests. Again, one of <strong>the</strong> key<br />

attributes of southwestern mixed-conifer is its<br />

inherent variability and diversity.<br />

At <strong>the</strong> warm/dry end of <strong>the</strong> environmental<br />

continuum, mixed-conifer <strong>for</strong>est typically<br />

intergrades with ponderosa pine <strong>for</strong>est. Where<br />

Douglas-fir, white fir, or blue spruce, ei<strong>the</strong>r<br />

singly or in combination, constitute less than<br />

5% cover or are considered “accidental” in late<br />

successional stands, <strong>the</strong>se stands are not included<br />

in <strong>the</strong> mixed-conifer <strong>for</strong>est classification.<br />

At <strong>the</strong> cold/wet end of <strong>the</strong> environmental<br />

continuum, mixed-conifer <strong>for</strong>est typically<br />

intergrades with subalpine spruce-fir <strong>for</strong>est.<br />

Where corkbark fir (Abies lasiocarpa var.<br />

arizonica), subalpine fir (A. lasiocarpa var.<br />

lasiocarpa), or Englemann spruce, ei<strong>the</strong>r singly<br />

or in common, constitute more than 5% of <strong>the</strong><br />

cover or are not considered “accidental,” <strong>the</strong><br />

<strong>for</strong>est is subalpine and no longer considered<br />

mixed-conifer.<br />

In addition to this general description,<br />

numerous authors have discussed aspects of


southwestern mixed-conifer <strong>for</strong>ests or classification<br />

of southwestern <strong>for</strong>ests. Pearson (1931)<br />

described a “Douglas-fir Zone.” He stated that<br />

although Douglas-fir is generally regarded as <strong>the</strong><br />

characteristic tree of this type, it rarely occurs in<br />

pure stands. Instead, it commonly occurs in<br />

stands with white fir, limber or <strong>Mexican</strong> white<br />

pine, and blue spruce. Western yellow pine (e.g.,<br />

ponderosa pine) is common in <strong>the</strong> lower portion<br />

of <strong>the</strong> type, and Engelmann spruce is common<br />

in <strong>the</strong> upper portion. Quaking aspen is common<br />

throughout <strong>the</strong> type.<br />

Choate (1966) also described Douglas-fir<br />

<strong>for</strong>ests in New Mexico as seldom growing in<br />

pure stands. He also stated that it mixes with<br />

ponderosa pine at lower elevations and with true<br />

firs and spruce at <strong>the</strong> upper limits. White fir and<br />

quaking aspen are common associates throughout<br />

<strong>the</strong> Douglas-fir type.<br />

USDA (1992) described old-growth attributes<br />

by cover types, including a “mixedspecies<br />

group” Forest Cover Type, which included<br />

<strong>the</strong> Douglas-fir, white fir, blue spruce,<br />

and limber pine <strong>for</strong>est cover types. They described<br />

<strong>the</strong>se mixed-species stands as having a<br />

rich diversity of vegetation, typically including at<br />

least three tree species.<br />

Moir (1993) described mixed-conifer <strong>for</strong>ests<br />

as upper montane coniferous <strong>for</strong>ests featuring<br />

Douglas-fir, white fir, several tall pine species,<br />

blue spruce, and quaking aspen. He included <strong>the</strong><br />

following series in this general <strong>for</strong>est type: Blue<br />

Spruce, White fir - Douglas-fir, Douglas-fir -<br />

Southwestern White Pine, White Fir - Douglasfir<br />

- Ponderosa Pine, Douglas-fir - Limber Pine -<br />

Bristlecone Pine, Douglas-fir - Gambel Oak, and<br />

Douglas-fir - Silverleaf Oak. These <strong>for</strong>ests are<br />

very productive because of ample precipitation<br />

and soils that are well watered throughout <strong>the</strong><br />

long growing season.<br />

Fletcher and Hollis (1994) described mixedconifer<br />

<strong>for</strong>est cover types as those dominated by<br />

Douglas-fir and/or white fir, usually containing<br />

varying amounts of ponderosa pine, southwestern<br />

white pine, and/or limber pine. Hardwood<br />

species, including rocky mountain maple (Acer<br />

glabrum), boxelder (A. negundo), bigtooth maple<br />

(A. grandidentatum), Gambel oak, quaking<br />

aspen, and o<strong>the</strong>r hardwood species may also be<br />

present. Douglas-fir and/or white fir typically<br />

Volume I/Part II<br />

54<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

comprise at least 40 percent and hardwood<br />

species less than 40 percent of <strong>the</strong> stand basal<br />

area. Conifers typical of higher elevations, such<br />

as Engelmann spruce, blue spruce, and/or subalpine<br />

fir may occur as “accidentals,” or provide<br />

less than about 5% cover in late successional<br />

stands.<br />

Layser and Schubert (1979), Moir and<br />

Ludwig (1979), Alexander et al. (1984a, b),<br />

Alexander and Ronco (1987), Youngblood and<br />

Mauk (1985), DeVelice et al. (1986), and<br />

Fitzhugh et al. (1987) all discussed classification<br />

of <strong>for</strong>est types in general or mixed-conifer <strong>for</strong>est<br />

types in particular in <strong>the</strong> southwestern United<br />

States. These treatments vary somewhat, possibly<br />

because of regional differences in <strong>for</strong>est types. A<br />

general consensus, however, indicates that<br />

mixed-conifer <strong>for</strong>est types generally fall in <strong>the</strong><br />

following four series: Abies concolor, Psuedotsuga<br />

menziesii, Pinus flexilis, or Picea pungens.<br />

Spruce-Fir<br />

Spruce-Fir<br />

Spruce-fir <strong>for</strong>ests in <strong>the</strong> southwestern United<br />

States generally coincide with <strong>the</strong> Subalpine<br />

Coniferous Forest discussed by Moir (1993).<br />

These are high-elevation <strong>for</strong>ests occurring on<br />

cold sites. They have short growing seasons,<br />

heavy snow accumulations, and strong ecological<br />

and floristic affinities to cold <strong>for</strong>ests of higher<br />

latitudes. Dominant trees include Engelmann<br />

spruce, subalpine and/or corkbark fir, or sometimes<br />

bristlecone pine. Moir and Ludwig (1979)<br />

included <strong>the</strong> Picea engelmannii and Abies<br />

lasiocarpa Series in <strong>the</strong> general spruce-fir <strong>for</strong>est<br />

type. Moir (1993) included <strong>the</strong> following series<br />

in this general <strong>for</strong>est group: Bristlecone pine,<br />

Engelmann spruce - bristlecone pine, corkbark<br />

fir - Engelmann spruce, Corkbark fir - Engelmann<br />

spruce - white fir, Engelmann spruce -<br />

Douglas-fir, and Engelmann spruce - limber<br />

pine.<br />

O<strong>the</strong>r O<strong>the</strong>r Forest Forest Types Types of of Interest<br />

Interest<br />

Chihuahua Chihuahua Pine<br />

Pine<br />

The Pinus leiophylla Series is described by<br />

Layser and Schubert (1979). This series typically


contains a diverse mixture of conifers and<br />

evergreen oaks. The conifer component is<br />

extensive enough to characterize this series as<br />

<strong>for</strong>est ra<strong>the</strong>r than woodland. Dominant conifers<br />

are typically Chihuahua pine, Apache pine, and<br />

P. ponderosa var. arizonica (Layser and Schubert<br />

1979). This series is Madrean in affinity and,<br />

within <strong>the</strong> U.S., is restricted to central and<br />

sou<strong>the</strong>rn Arizona and southwestern New Mexico<br />

(Brown et al. 1980). Moir (1993) included this<br />

type in <strong>the</strong> Lower Montane Coniferous Forest<br />

group.<br />

Quaking Quaking Quaking Aspen<br />

Aspen<br />

Quaking aspen is a special feature of western<br />

landscapes. It is a major seral species in <strong>the</strong><br />

following series; Abies lasiocarpa, Picea pungens,<br />

and Abies concolor. It is a minor seral species in<br />

<strong>the</strong> Picea engelmannii, Pseudotsuga menziesii, and<br />

Pinus ponderosa Series (Larson and Moir 1986).<br />

As such, quaking aspen should be a common<br />

component of <strong>the</strong>se landscapes under natural<br />

disturbance regimes.<br />

Riparian Riparian Forests<br />

Forests<br />

Numerous authors have discussed classification<br />

and ecology of riparian <strong>for</strong>ests in <strong>the</strong> southwestern<br />

United States (e.g., Pase and Layser<br />

1977, Layser and Schubert 1979, Brown et al.<br />

1980, Medina 1986, Szaro 1989). In general,<br />

southwestern riparian <strong>for</strong>ests are dominated by<br />

various species of broadleaved deciduous trees<br />

and shrubs. Trees common in adjacent uplands,<br />

such as conifers, oaks, and quaking aspen, may<br />

occur in association with riparian trees, but<br />

generally do not dominate <strong>the</strong> site (Brown<br />

1982).<br />

PLAN PLAN DEFINITIONS<br />

DEFINITIONS<br />

Our classification scheme is primarily<br />

concerned with a subset of <strong>the</strong> available <strong>for</strong>est<br />

types in <strong>the</strong> southwestern United States. We are<br />

interested in both potential and existing vegetation.<br />

Consequently, our scheme is a hybrid of<br />

classification schemes based on potential vegetation<br />

(series, association and habitat type) and<br />

<strong>for</strong>est cover types based on existing vegetation.<br />

Volume I/Part II<br />

55<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Three terms used in our definitions require<br />

clarification here. These are “pure,” “majority,”<br />

and “plurality.” Various definitions exist to<br />

describe what constitutes a pure stand. Daniel et<br />

al. (1979) described pure stands as those where<br />

>90% of <strong>the</strong> dominant or codominant trees are<br />

of a single species. A stand may have an understory<br />

of o<strong>the</strong>r species without changing <strong>the</strong> pure<br />

designation. The key to this concept is <strong>the</strong><br />

distinction between <strong>the</strong> dominant and codominant<br />

species and <strong>the</strong> understory component. In<br />

contrast, Eyre (1980) defined a pure stand as one<br />

where >80% of <strong>the</strong> stocking is by one species.<br />

For purposes of this plan, we use <strong>the</strong> term<br />

pure to refer to any stand where a single species<br />

contributes >80% of <strong>the</strong> basal area of dominant<br />

and codominant trees. We use <strong>the</strong> term majority<br />

to refer to <strong>the</strong> situation where a single species<br />

contributes >50% of <strong>the</strong> basal area (Eyre 1980).<br />

We use <strong>the</strong> term plurality to refer to <strong>the</strong> situation<br />

where a species (or group of species of<br />

interest) comprises <strong>the</strong> largest proportion of a<br />

mixed-species stand (Eyre 1980).<br />

With <strong>the</strong>se definitions and concepts in<br />

mind, definitions <strong>for</strong> specific <strong>for</strong>est cover types<br />

are provided below.<br />

Ponderosa Ponderosa Pine Pine Forest<br />

Forest<br />

as:<br />

We define <strong>the</strong> Ponderosa Pine Forest Type<br />

1. Any <strong>for</strong>ested stand of <strong>the</strong> Pinus ponderosa<br />

Series not included in <strong>the</strong> Pine-Oak<br />

Forest Type (see below), or;<br />

2. Any stand that qualifies as pure (Eyre<br />

1980) ponderosa pine, regardless of <strong>the</strong><br />

series or habitat type.<br />

Pine-oak Pine-oak Forest<br />

Forest<br />

A number of habitat types exist in <strong>the</strong><br />

southwestern United States that could be described<br />

as pine-oak. Most of <strong>the</strong> stands relevant<br />

to recovery of <strong>the</strong> <strong>Mexican</strong> spotted owl fall<br />

within two series, <strong>the</strong> Pinus ponderosa Series and<br />

<strong>the</strong> Pinus leiophylla Series. Present evidence,<br />

however, suggests that <strong>the</strong> <strong>for</strong>mer series includes


many areas that could never attain <strong>the</strong> type of<br />

<strong>for</strong>est structure sought by spotted owls <strong>for</strong><br />

roosting and nesting. There<strong>for</strong>e, in an attempt to<br />

avoid needlessly restricting management options<br />

on lands not used to any great extent by <strong>the</strong><br />

spotted owl, we propose <strong>the</strong> following operational<br />

definition <strong>for</strong> pine-oak <strong>for</strong>est under this<br />

<strong>Plan</strong>:<br />

1. Any stand within <strong>the</strong> Pinus leiophylla<br />

Series.<br />

2. Any stand within <strong>the</strong> Pinus ponderosa<br />

Series that meets <strong>the</strong> following criteria<br />

simultaneously:<br />

Volume I/Part II<br />

a) Habitat types that reflect Quercus<br />

gambelii or a Quercus gambelii phase<br />

of <strong>the</strong> habitat type.<br />

b) The stand is located in ei<strong>the</strong>r <strong>the</strong><br />

Upper Gila Mountains <strong>Recovery</strong><br />

Unit, <strong>the</strong> Basin and Range-West<br />

<strong>Recovery</strong> Unit, or <strong>the</strong> Zuni Mountains<br />

or Mount Taylor regions of <strong>the</strong><br />

Colorado Plateau <strong>Recovery</strong> Unit.<br />

c) >10% of <strong>the</strong> stand basal area or<br />

2.3 m 2 /ha (10 ft 2 /ac) of basal area<br />

consists of Gambel oak > 13 cm<br />

(5 in) diameter at root collar.<br />

3. Any stand within <strong>the</strong> Basin and Range-<br />

West <strong>Recovery</strong> Unit of any o<strong>the</strong>r series<br />

that meets <strong>the</strong> following criteria<br />

simultaneously:<br />

a) A plurality (Eyre 1980) of <strong>the</strong> basal<br />

area exists in yellow pines (ponderosa,<br />

Arizona, Apache, or Chihuahua).<br />

b) >10% of <strong>the</strong> stand basal area or<br />

2.3 m 2 /ha (10 ft 2 /ac) of basal area<br />

consists of any oaks > 13 cm (5 in)<br />

diameter at root collar.<br />

56<br />

Mixed-conifer Mixed-conifer Forest<br />

Forest<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Natural variability is high within this <strong>for</strong>est<br />

type and has been increased by both natural and<br />

human-caused disturbances. Despite this variability,<br />

an extant classification scheme based on<br />

series and habitat types (Hanks et al. 1983,<br />

Layser and Schubert 1979, Alexander et al. 1984<br />

a, b; Alexander and Ronco 1987, Youngblood<br />

and Mauk 1985, DeVelice et al. 1986, Fitzhugh<br />

et al. 1987) is available. This classification system<br />

is in widespread use and has multiple-agency<br />

support. Given that background, we propose<br />

using that system as a starting point in defining<br />

mixed-conifer <strong>for</strong>est, with some added refinements.<br />

Specifically, we propose that:<br />

1. The definition of mixed-conifer <strong>for</strong>est<br />

generally be confined to <strong>the</strong> following<br />

series (Layser and Schubert 1979) and<br />

associated habitat types (after authors<br />

listed above): Abies concolor, Pseudotsuga<br />

menziesii, Pinus flexilis, or Picea pungens.<br />

Within this framework, we provide <strong>the</strong><br />

following exceptions to <strong>the</strong> general guideline<br />

listed above:<br />

1. Any stand within <strong>the</strong> Pinus aristata,<br />

Picea engelmannii, or Abies lasiocarpa<br />

Series not having a plurality (Eyre 1980)<br />

of basal area of any of Pinus aristata,<br />

Picea engelmannii, Abies lasiocarpa, or<br />

Pinus ponderosa, singly or in combination,<br />

should also be defined as mixedconifer.<br />

2. Stands that can be described as “pure” <strong>for</strong><br />

coniferous species o<strong>the</strong>r than Douglasfir,<br />

white fir, southwestern white pine,<br />

limber pine, or blue spruce should be<br />

excluded from <strong>the</strong> broad category of<br />

mixed-conifer <strong>for</strong> <strong>the</strong> purposes of <strong>Plan</strong><br />

implementation regardless of <strong>the</strong> series<br />

or habitat type. By pure, we mean that<br />

one species comprises 80% or more of<br />

<strong>the</strong> dominant and codominant trees<br />

(Eyre 1980).


3. Stands of mixed species with >50% of<br />

<strong>the</strong> basal area consisting of quaking<br />

aspen should be defined as quaking<br />

aspen <strong>for</strong> <strong>the</strong> purposes of <strong>Plan</strong> implementation<br />

regardless of <strong>the</strong> series or<br />

habitat type.<br />

High-elevation High-elevation Forests,<br />

Forests,<br />

Including Including Spruce-fir Spruce-fir Forest<br />

Forest<br />

We define this <strong>for</strong>est type as:<br />

1. Any stand of <strong>the</strong> Pinus aristata, Picea<br />

engelmannii, or Abies lasiocarpa Series<br />

that meets <strong>the</strong> following criteria:<br />

Volume I/Part II<br />

a) The majority (Eyre 1980) of stand<br />

basal area consists of any of <strong>the</strong> three<br />

species listed above, ei<strong>the</strong>r singly or<br />

in combination, or;<br />

b) Any stand that qualifies as a pure<br />

stand (Eyre 1980) of any of <strong>the</strong>se<br />

species, regardless of <strong>the</strong> series or<br />

habitat type.<br />

Quaking Quaking Quaking Aspen<br />

Aspen<br />

We propose that any stands with >50% of<br />

<strong>the</strong> basal area consisting of quaking aspen be<br />

defined as quaking aspen.<br />

KEY KEY TO TO FOREST FOREST COVER COVER TYPES<br />

TYPES<br />

1. Trees deciduous and broadleaved, often<br />

confined to floodplain, drainageway, or<br />

canyon bottom (Layser and Schubert<br />

1979) . . . . . . . . . . . . . . Ripar Riparian Ripar ian F FFor<br />

F For<br />

or orest or est<br />

1. Dominant trees evergreen and needleleaved<br />

.. . . . . . . . . . . . . . . . . . . . . . . . ..2<br />

2. Series = Psuedotsuga menziesii, Abies<br />

concolor, Pinus flexilis, or Picea<br />

pungens . . . . . . . . . . . . . . . . . . . . . .3<br />

2. Series not as above . . . . . . . . . . . . .5<br />

57<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

3. >80% of dominant and codominant<br />

trees are species o<strong>the</strong>r than Pseudotsuga<br />

menziesii, Abies concolor, Pinus<br />

strobi<strong>for</strong>mis, Pinus flexilis, or Picea<br />

pungens . . . . . . . . Classify Classify bb<br />

by bb<br />

y dominant<br />

dominant<br />

species<br />

species<br />

3. Stand not as above . . . . . . . . . . . . . . . .4<br />

4. Populus tremuloides contributes<br />

>50% of stand basal area<br />

. . . . . . . . . . Quaking uaking Aspen Aspen F FFor<br />

F or orest or est<br />

4. Not as above. . . . . . . Mix ix ixed-conifer ix ed-conifer<br />

For or orest or est<br />

5. Series = Pinus leiophylla . . . . . Pine-oak ine-oak<br />

For or or orest orest<br />

est<br />

5. Series not as above . . . . . . . . . . . . . . . . 6<br />

6. Series = Pinus ponderosa . . . . . . . . . 7<br />

6. Series not as above . . . . . . . . . . . . 10<br />

7. Habitat type or phase includes Quercus<br />

gambelii . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

7. Not as above. . . . . . . . . . . . . .Ponder onder onderosa onder osa<br />

Pine ine F FFor<br />

F or orest or est<br />

8. Area is located within Upper Gila<br />

Mountains <strong>Recovery</strong> Unit, Basin and<br />

Range-West <strong>Recovery</strong> Unit, or <strong>the</strong><br />

sou<strong>the</strong>astern portion of <strong>the</strong> Colorado<br />

Plateau <strong>Recovery</strong> Unit (Zuni Mtns.,<br />

Mt. Taylor). . . . . . . . . . . . . . . . . . . 9<br />

8. Area not located as above. . . . . . . . . .<br />

. . . .. . . . . . . Ponder onder onder onderosa onder osa P PPine<br />

P ine F FFor<br />

F or orest or est<br />

9. >10% of stand basal area or 2.3 m2 /ha<br />

(10 ft2 /ac) consists of Quercus gambelii ><br />

13 cm (5 in) diameter at root collar. . . .<br />

Pine-oak ine-oak F FFor<br />

F or orest or est<br />

9. Not as above . . . . . . . . . . . . . Ponder onder onderosa onder osa<br />

Pine ine F FFor<br />

F or orest orest<br />

est


Volume I/Part II<br />

10. Series = Pinus aristata, Picea<br />

engelmannii, or Abies lasiocarpa<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

10. Series not as above . . . . . . . . . . . 13<br />

11. Stand can be defined as pure <strong>for</strong> ei<strong>the</strong>r<br />

Pinus aristata, Picea engelmannii, or Abies<br />

lasiocarpa . . . . . . . . . . SS<br />

Spr S pr pruce-fir pr uce-fir F FFor<br />

F or orest or est<br />

11. Stand not as above . . . . . . . . . . . . . . . 12<br />

12. Pinus aristata, Picea engelmannii,<br />

or Abies lasiocarpa contribute >50%<br />

of stand basal area, ei<strong>the</strong>r singly or<br />

in combination. . . . . . . S SSpr<br />

S pr pruce-fir pr uce-fir<br />

For or orest or est<br />

12. Stand not as above . . . . . . . . . . . . .<br />

. . . . . . . . . . .Mix ix ixed-conifer ix ed-conifer F FFor<br />

F or orest or est<br />

58<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

13. Stand located in Basin and Range-West<br />

<strong>Recovery</strong> Unit . . . . . . . . . . . . . . . . . . 14<br />

13. Stand not located as above . . . . . . O<strong>the</strong>r <strong>the</strong>r<br />

14. A plurality of stand basal area is<br />

contributed by Pinus ponderosa,<br />

Pinus engelmannii, or Pinus<br />

leiophylla, ei<strong>the</strong>r singly or in<br />

combination . . . . . . . . . . . . . . . . .15<br />

14. Stand not as above . . . . . . . . . O<strong>the</strong>r <strong>the</strong>r<br />

15. >10% of stand basal area or 2.3 m2 /ha<br />

(10ft2 ac) consists of any oak > 13 cm (5<br />

in) diameter at root collar. . . . . . . . . . . .<br />

. .. . . . . . . . . . . . . . . . . PP<br />

Pine-oak P ine-oak F FFor<br />

F For<br />

or orest or est<br />

15. Stand not as above . . . . . . . . . . . . O<strong>the</strong>r <strong>the</strong>r


The goal of this <strong>Recovery</strong> <strong>Plan</strong> is to recover<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl so that it no longer<br />

requires protection under <strong>the</strong> Endangered<br />

Species Act. We submit that this can be best<br />

achieved by ensuring a mosaic of all successional<br />

stages, now and in <strong>the</strong> future, throughout a<br />

landscape comprised of all known habitat types<br />

used by <strong>the</strong> owl. In mixed-conifer, pine-oak and<br />

riparian <strong>for</strong>ests, this habitat mosaic must contain<br />

stands adequate <strong>for</strong> all life-history requirements,<br />

including nesting and roosting.<br />

We agree with <strong>the</strong> widely held belief that<br />

conditions within some southwestern <strong>for</strong>ests<br />

deviate substantially from those existing prior to<br />

European settlement. Moreover, <strong>for</strong>ests throughout<br />

<strong>the</strong> U.S. range of <strong>the</strong> owl are at high risk<br />

from fire, insects, and disease. The mechanisms<br />

responsible <strong>for</strong> current condition are not completely<br />

known, but synergistic effects of past<br />

timber harvest, overgrazing, and fire suppression<br />

are plausible explanations. The intent of this<br />

<strong>Recovery</strong> <strong>Plan</strong> is not to cast blame on any<br />

particular aspect of past management, but to<br />

outline <strong>the</strong> appropriate steps needed to ensure<br />

persistence of <strong>the</strong> <strong>Mexican</strong> spotted owl. Thus,<br />

<strong>the</strong> basis to maintain owl populations is to<br />

ensure that adequate habitat quality and quantity<br />

will be sustained through time. These conditions<br />

also must be within <strong>the</strong> natural range of variation.<br />

We recognize, however, that major knowledge<br />

gaps preclude accurate descriptions of <strong>the</strong><br />

natural range of variation and presettlement<br />

conditions. We cannot verify that fewer owls<br />

exist today than 100 years ago, or vice versa. We<br />

know little about habitat quality and how<br />

contemporary landscapes and ecosystem conditions<br />

contribute to owl fitness and population<br />

persistence. Thus, management of <strong>the</strong> owl<br />

should proceed in an iterative fashion. We must<br />

use <strong>the</strong> best available knowledge to guide current<br />

management, recognizing that new in<strong>for</strong>mation<br />

from research and monitoring is critical <strong>for</strong> <strong>the</strong><br />

development of long-term management plans.<br />

Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

D. D. CONCEPTUAL CONCEPTUAL FRAMEWORK FRAMEWORK FOR FOR RECOVERY<br />

RECOVERY<br />

William H. Moir, James L. Dick Jr., William M.<br />

Block, James P. Ward Jr., Robert Vahle,<br />

Frank P. Howe, and Joseph L. Ganey<br />

59<br />

This <strong>Recovery</strong> <strong>Plan</strong> details a short-term (10-<br />

15 years) strategy aimed at maintaining owl<br />

habitat where it exists and initiating a process to<br />

develop a <strong>for</strong>ested landscape that includes<br />

replacement habitat. We presently have insufficient<br />

knowledge to design a strategy that will<br />

answer all long-term considerations, such as<br />

allocation of stand structures in space and time.<br />

Under proposed delisting criteria (III.A), <strong>the</strong> owl<br />

could be delisted within this 10-15 years, rendering<br />

this plan obsolete.<br />

To achieve <strong>the</strong> recovery goals outlined in this<br />

<strong>Plan</strong>, management must emulate natural ecosystem<br />

processes and landscape mosaics that balance<br />

natural variability and secure <strong>the</strong> landscape<br />

against catastrophic habitat loss. Our recommendations<br />

assume that population status and<br />

habitat condition will be monitored in conjunction<br />

with recovery ef<strong>for</strong>ts <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl (Part III). The management recommendations<br />

are not meant to stand alone without<br />

such monitoring.<br />

ECOSYSTEM ECOSYSTEM OR OR LANDSCAPE<br />

LANDSCAPE<br />

MANAGEMENT<br />

MANAGEMENT<br />

Volume 2 summarizes current knowledge of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl’s basic natural and life<br />

histories, but a brief reiteration is appropriate<br />

here. First, <strong>the</strong> owl is found in a number of<br />

different habitat types ranging from slickrock<br />

canyons to cool, mesic <strong>for</strong>ests. Second, <strong>the</strong> owl<br />

has relatively large home ranges, typically containing<br />

mosaics of vegetation types and different<br />

seral stages and conditions within those types.<br />

Third, <strong>the</strong> owl takes numerous species of prey<br />

and each of <strong>the</strong>se species has unique habitat<br />

requirements. These factors considered simultaneously<br />

stress <strong>the</strong> need to consider management<br />

across spatial scales ranging from sites to landscapes<br />

and to provide <strong>the</strong> diversity of conditions<br />

required <strong>for</strong> <strong>the</strong> owl’s life history. Consequently,<br />

we submit that management <strong>for</strong> <strong>Mexican</strong> spotted<br />

owls must be viewed within <strong>the</strong> context of<br />

managing ecosystems.


How can ecosystem management be applied<br />

to <strong>the</strong> <strong>Mexican</strong> spotted owl? Ecosystem management<br />

should sustain biotic diversity and <strong>the</strong><br />

natural processes and landscape mosaics that<br />

generate that diversity (cf., Jensen and<br />

Bourgeron 1993, Franklin 1993, 1994; Diaz<br />

and Apostol 1993, Kaufmann et al. 1994,<br />

Williams 1994). The current emphasis in<br />

ecosystem management is to use <strong>the</strong> filter<br />

approach described by Hunter (1991). Two<br />

“sizes” of filters, coarse and fine, are used.<br />

The objective of <strong>the</strong> coarse filter approach is<br />

to maintain <strong>the</strong> natural array of conditions that<br />

exist within <strong>the</strong> biotic and physical limits of <strong>the</strong><br />

landscape. This would include special as well as<br />

common habitats. Ideally, <strong>the</strong> array of conditions<br />

provided by using <strong>the</strong> coarse-filter approach<br />

should maintain most plants and animals<br />

adapted to natural conditions (Hunter 1991).<br />

This should include most of <strong>the</strong> habitat conditions<br />

needed by <strong>the</strong> owl and its prey.<br />

In some cases, however, a fine filter may be<br />

required <strong>for</strong> specific habitats, or habitat elements,<br />

that fall through <strong>the</strong> coarse filter. With<br />

respect to <strong>the</strong> <strong>Mexican</strong> spotted owl, <strong>the</strong> coarse<br />

filter is probably sufficient <strong>for</strong> most <strong>for</strong>aging<br />

habitats, but a fine filter may be needed to<br />

provide nest and roost sites. For example, <strong>the</strong><br />

owl prefers or needs particular land<strong>for</strong>ms (such<br />

as steep-walled canyons), particular structures<br />

(including snags, large trees, large logs, cavities<br />

and o<strong>the</strong>r plat<strong>for</strong>ms <strong>for</strong> nesting), mature <strong>for</strong>ests,<br />

and specialized microhabitats. Thus, a fine-filter<br />

analysis is required to identify and ensure continuing<br />

availability of <strong>the</strong> owl’s specific habitat<br />

needs.<br />

In summary, <strong>the</strong> coarse filter approach is<br />

used to manage <strong>the</strong> overall landscape, and, if<br />

properly applied, should suffice to maintain <strong>the</strong><br />

natural array of conditions on that landscape.<br />

The fine filter is used to provide specialized<br />

habitats or habitat elements within that overall<br />

landscape.<br />

Two <strong>the</strong>mes of <strong>the</strong> recovery measures are<br />

consistent with <strong>the</strong>se principles of ecosystem<br />

management. The first <strong>the</strong>me is that <strong>the</strong> general<br />

recommendations of this <strong>Recovery</strong> <strong>Plan</strong> provide<br />

conditions <strong>for</strong> <strong>the</strong> owl across <strong>the</strong> landscape. This<br />

landscape should provide nesting, roosting,<br />

<strong>for</strong>aging, and dispersal macrohabitats in <strong>the</strong><br />

Volume I/Part II<br />

60<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

short term. This <strong>the</strong>me emphasizes protecting<br />

and monitoring owl populations and habitats.<br />

The second <strong>the</strong>me acknowledges that ecosystems<br />

are temporally dynamic, and that provisions<br />

are needed to ensure owl habitat in <strong>the</strong><br />

long term. As nest sites change and are abandoned,<br />

new nest sites should develop and become<br />

occupied. Allocation of mid- to late-seral<br />

<strong>for</strong>ests needed by spotted owls, and o<strong>the</strong>r species,<br />

in future decades requires knowledge of<br />

<strong>for</strong>est disturbances, risks, and rates of succession<br />

at different spatio-temporal scales. We outline<br />

below some disturbances, risks, and tools that<br />

should be considered in managing present and<br />

future owl habitat.<br />

Fire<br />

Fire<br />

Fire is <strong>the</strong> most rapidly acting of natural<br />

disturbances. A crown fire can quickly consume<br />

<strong>for</strong>ests across vast tracts. After a large crown fire,<br />

habitat components <strong>for</strong> nesting, roosting, and<br />

<strong>for</strong>aging are reduced or eliminated. Small-scale<br />

natural fires and prescribed burns, however, can<br />

reduce fuel loadings and create small openings<br />

and thinned stands that increase horizontal<br />

diversity and reduce <strong>the</strong> spread of catastrophic<br />

fire. Small-scale fires and lightning also create<br />

snags, canopy gaps, and large logs, plus <strong>the</strong>y<br />

perpetuate understory shrubs, grasses, and <strong>for</strong>bs<br />

which are important habitat components to <strong>the</strong><br />

owl, its prey, and o<strong>the</strong>r wildlife. Under natural<br />

fire regimes prior to 1890 <strong>the</strong>se small fires<br />

occurred frequently (Moody et al. 1992).<br />

The risk of catastrophic fires is widespread in<br />

Southwestern <strong>for</strong>ests and woodlands (Moody et<br />

al 1992). Fuel accumulations and <strong>for</strong>ests overstocked<br />

with trees place spotted owl habitat at<br />

risk with respect to stand-replacing fires. Figures<br />

II.D.1-3 show <strong>the</strong> changing fire record from<br />

1910 to 1992, based on records compiled at <strong>the</strong><br />

FS Southwestern Regional Office. Because FS<br />

burn policies changed during this period and <strong>the</strong><br />

use of prescribed natural fire increased, interpretation<br />

of <strong>the</strong>se records is not straight<strong>for</strong>ward. In<br />

general, however, <strong>the</strong> figures document an<br />

increase in both area burned per year and in area<br />

lost to catastrophic, stand-replacing fires.<br />

The number of total natural and humancaused<br />

fires generally declined after 1981 (Figure


II.D.1), but <strong>the</strong> number of large fires (> 4 ha<br />

[>10 acres]) increased at <strong>the</strong> same time (Figure<br />

II.D.2). Figure II.D.3 shows <strong>the</strong> trend clearly;<br />

from 1985 to 1992 <strong>the</strong> number of hectares that<br />

burned increased. If <strong>the</strong> influence of two exceptional<br />

fire years (1974 and 1979) is removed, <strong>the</strong><br />

trend shown in Figure II.D.3 remains; that is,<br />

<strong>the</strong> number of large fires increased.<br />

Moody et al. (1992) estimated that about<br />

303,500 ha [750,000 acres] of mixed-conifer<br />

<strong>for</strong>est within FS Region 3 needed treatment to<br />

reduce fire risk in <strong>the</strong> next 10 years. Unmanaged<br />

and unplanned conversion of large areas of<br />

<strong>for</strong>ests or woodlands to early seral conditions by<br />

wildfire can disrupt management goals to maintain<br />

existing and to provide future spotted owl<br />

habitat (USDA 1993c).<br />

Characteristics of many nest and roost sites<br />

of spotted owls place <strong>the</strong>m at high fire risk.<br />

Some nest/roost locations at special topographic<br />

locations (such as steep-walled canyons or<br />

isolated places) may be fire refugia, however.<br />

Taking cue from <strong>the</strong>se, one promising management<br />

tactic is to isolate nest/roost sites from <strong>the</strong><br />

adjoining high-risk <strong>for</strong>est by reducing flammability<br />

and fire spread in a buffer around <strong>the</strong> site.<br />

This must be done, of course, without compromising<br />

<strong>the</strong> site itself as nest/roost habitat.<br />

Inevitably, severe climatic conditions will<br />

occur in <strong>the</strong> future, and extreme fire years are<br />

possible (Swetnam and Betancourt 1990). Given<br />

<strong>the</strong> present conditions of Southwestern <strong>for</strong>ests,<br />

extreme fire years could result in holocaustic fires<br />

throughout large portions of <strong>the</strong> owl’s range.<br />

Because <strong>the</strong> resulting damage to owl habitat<br />

would be irreparable in <strong>the</strong> <strong>for</strong>seeable future,<br />

ef<strong>for</strong>ts to limit large-scale catastrophic fires are of<br />

utmost importance <strong>for</strong> owl conservation.<br />

Increased use of fire and o<strong>the</strong>r tools will be<br />

needed to reduce <strong>the</strong> amount of <strong>for</strong>est at high<br />

risk from stand-replacing fires. The <strong>Recovery</strong><br />

Team encourages proactive fire management<br />

programs which assume active roles in fuels<br />

management and understanding <strong>the</strong> ecological<br />

role of fire. An example of such a program is <strong>the</strong><br />

one employed by <strong>the</strong> Gila National Forest.<br />

The <strong>Recovery</strong> Team recognizes that fire<br />

technology may not be at <strong>the</strong> level of sophistication<br />

needed to maintain owl habitat and create<br />

new habitat. Although we advocate broadscale<br />

Volume I/Part II<br />

61<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

use of fire in <strong>the</strong> Southwest, we also stress <strong>the</strong><br />

need to approach <strong>the</strong> use of fire in an adaptive<br />

management context. Prescriptions that maintain<br />

key structural features of owl and small prey<br />

habitats should be developed and tested. These<br />

features include large trees (which are often fire<br />

resistant), snags, logs, and understory hardwood<br />

trees. Treatments to produce or maintain such<br />

habitat components must be assessed by monitoring<br />

to evaluate if treatment objectives were<br />

met in both short and long terms. Wholesale use<br />

of fire without understanding or monitoring its<br />

effects on habitat may render areas unusable by<br />

owls, and may also miss opportunities to improve<br />

our knowledge of fire effects. Fire and<br />

wildlife personnel should work toge<strong>the</strong>r to refine<br />

fire prescriptions compatible with maintenance<br />

of important habitat elements.<br />

O<strong>the</strong>r O<strong>the</strong>r Natural Natural Disturbances<br />

Disturbances<br />

Disturbances<br />

The vegetative communities that provide<br />

habitat <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl are dynamic<br />

assemblages of living plants, snags, logs, and<br />

numerous organisms active in decay and nutrient-cycling<br />

processes. Herbivory, disease, and<br />

structural change caused by bacteria, fungi,<br />

insects, and vertebrates are natural agents of<br />

change in <strong>for</strong>est and woodland communities and<br />

occur at scales ranging from individual trees to<br />

landscapes.<br />

These disturbances contribute to <strong>the</strong> <strong>for</strong>mation<br />

of complex landscape mosaics in which<br />

woodlands and <strong>for</strong>ests consist of aggregates of<br />

transient patches and gaps. Added to this patchiness<br />

are changes of <strong>the</strong> structural elements of owl<br />

habitat caused by disturbances at scales larger<br />

than gaps. Climate change, pollutants, and o<strong>the</strong>r<br />

extensive events will produce effects of magnitudes<br />

that are poorly understood (Davis 1989).<br />

Although management scarcely influences <strong>the</strong><br />

primary determinants of vegetation pattern<br />

(geology, climate, and genetics), management<br />

can affect vegetation by manipulating <strong>the</strong> extent,<br />

severity, and frequency of disturbance.<br />

Land managers should recognize that natural<br />

disturbances can create and maintain diverse and<br />

productive ecosystems that always include,<br />

somewhere on <strong>the</strong> landscape, an adequate<br />

amount and distribution of <strong>the</strong> vegetative


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e II.D.1. II.D.1. Historical record of number of total fires and number of natural fires. Data from<br />

USFS Southwestern Region.<br />

Figur igur igure igur e II.D.2. II.D.2. Historical record of fires over four hectares in size. Shown are numbers of fires and<br />

area burned. Data from USFS Southwestern Region.<br />

62


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e II.D.3. II.D.3. Five-year running averages of area burned. Data from USFS Southwestern Region<br />

elements that are <strong>the</strong> required habitat <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. The word “adequate” is<br />

crucial. Adequacy is derived from publicly<br />

acceptable landscape descriptions (desired<br />

conditions), toge<strong>the</strong>r with use of <strong>the</strong> best succession,<br />

allocation, and landscape-dynamic models<br />

to guide managers in how to get <strong>the</strong>re. Adequacy<br />

is tested by ongoing monitoring and adaptive<br />

management (III.C) and should not be assumed<br />

in <strong>the</strong> absence of monitoring.<br />

Insects and microorganisms can be beneficial<br />

as well as destructive agents of plant succession<br />

(Dinoor and Eshed 1984, Knauer 1988,<br />

Dickman 1992, Haack and Byler 1993). These<br />

organisms may produce large-scale community<br />

changes after periods of climatic stress that<br />

“predispose” <strong>for</strong>ests to insects or pathogenic<br />

occurrences (Colhoun 1979). Several groups of<br />

<strong>for</strong>est insects occasionally develop epidemic<br />

populations that severely damage mature <strong>for</strong>est<br />

trees over large areas. Among <strong>the</strong> defoliating<br />

insects, <strong>the</strong> western spruce budworm kills<br />

understory white fir and Douglas-fir and thins<br />

<strong>the</strong> crowns of overstory trees (Archambault et al.<br />

1994). Outbreaks of western spruce budworm<br />

63<br />

occur every decade or so and extend widely<br />

across <strong>the</strong> landscape. Perhaps as a response to fire<br />

exclusion policies, recent budworm outbreaks<br />

have tended to be regionally synchronous with<br />

<strong>the</strong> maturing of host species over large areas<br />

(Swetnam and Lynch 1993). As a complicating<br />

factor, trees that suffer declining vigor from<br />

multiple years of defoliation by budworms may<br />

lose <strong>the</strong>ir resistance to more injurious woodboring<br />

insects and ultimately die. Bark beetles<br />

are important wood-boring insects in pinyon,<br />

ponderosa pine, Douglas-fir, and Englemann<br />

spruce. During outbreaks (about every 7 to 10<br />

years), <strong>the</strong>se insects kill groups of mature trees.<br />

In longer outbreaks (usually those following<br />

droughts), mortality groups coalesce and damage<br />

appears to be widespread. Bark beetle populations<br />

are most likely to increase where host trees<br />

are stressed as a result of sublethal fire damage,<br />

dwarf mistletoe infection, or where abundant<br />

green slash is available from thinning or<br />

blowdown.<br />

The principal <strong>for</strong>est pathogens are root<br />

disease fungi and dwarf mistletoe. Armillaria<br />

root disease is widespread across <strong>the</strong> <strong>for</strong>ests of


<strong>the</strong> Southwest. In a few locations it behaves as an<br />

aggressive killing agent (Marsden et al. 1993),<br />

but in most stands it acts to remove trees weakened<br />

by lightning or insects. O<strong>the</strong>r root diseases<br />

are caused by Heterobasidion annosum and<br />

Phellinus schweinitzii.<br />

The most common tree disease in Southwestern<br />

<strong>for</strong>ests is caused by parasitic seed plants<br />

of <strong>the</strong> genus Arceuthobium, <strong>the</strong> dwarf mistletoes.<br />

About one-half to two-thirds of <strong>the</strong> stands in<br />

<strong>the</strong>se <strong>for</strong>ests are infested by dwarf mistletoe.<br />

Infected trees become stunted, develop witches’<br />

brooms, and are eventually killed by this or<br />

o<strong>the</strong>r mortality agents. Both root disease and<br />

mistletoe typically occur as “centers” or “patches”<br />

and create slowly but continously expanding<br />

canopy gaps. These agents increase ecosystem<br />

diversity by producing snags, logs, and, in <strong>the</strong><br />

case of mistletoe, witches brooms. They also act<br />

synergistically with <strong>for</strong>est insects.<br />

The relationship between fire and dwarf<br />

mistletoe is complex. Brooms caused by dwarf<br />

mistletoe provide fuel continuity from ground to<br />

tree crown. By maintaining seral trees in <strong>for</strong>est<br />

stands, fire increases <strong>the</strong> opportunity <strong>for</strong> mistletoe<br />

infection because <strong>the</strong> seral trees are more<br />

commonly hosts than climax trees. Similar<br />

complex relationships exist between fire, bark<br />

beetles, and western spruce budworms.<br />

White pine blister rust is caused by an exotic<br />

fungus that was recently introduced into <strong>the</strong><br />

Sacramento Mountains. It has <strong>the</strong> potential to<br />

kill most of <strong>the</strong> southwestern white pine in <strong>the</strong><br />

mixed-conifer <strong>for</strong>ests (Hawksworth and Conklin<br />

1990) where <strong>the</strong> greatest concentration of<br />

<strong>Mexican</strong> spotted owls occurs. Although southwestern<br />

white pine is seldom <strong>the</strong> most frequent<br />

tree species of a stand, it is an important seral,<br />

dominant, or codominant species in most areas.<br />

This tree produces large seeds and readily fills<br />

gaps opened by mortality of o<strong>the</strong>r trees to<br />

budworm, bark beetles, root disease, and mistletoe.<br />

There<strong>for</strong>e, <strong>the</strong> short-term effects of white<br />

pine blister rust may be negative, since a strong<br />

reordering of <strong>for</strong>est tree composition may take<br />

place. A number of actions can be taken to<br />

“control” <strong>the</strong> rust and reduce its impacts, but<br />

<strong>the</strong>y are expensive and <strong>the</strong>ir effectiveness and<br />

possible side effects are unknown. In <strong>the</strong> longterm,<br />

a genetic balance between <strong>the</strong> rust and<br />

Volume I/Part II<br />

64<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

white pines may occur, as it did with Pinus<br />

monticola in <strong>the</strong> nor<strong>the</strong>rn Rockies (Ledig 1992).<br />

Various o<strong>the</strong>r arthropods and saprophytic<br />

fungi are also important agents of deterioration<br />

and decay of snags and logs. Although <strong>the</strong>se<br />

agents generally do not kill trees directly, <strong>the</strong>ir<br />

activity (decay) can lead to stem breakage and<br />

tree death. They are thus important in determining<br />

<strong>the</strong> condition and persistence of coarse<br />

woody debris within <strong>for</strong>est stands.<br />

The cumulative impacts of <strong>the</strong>se disturbance<br />

agents on owl habitat depends on a number of<br />

factors, some of which are subject to manipulation.<br />

In general, <strong>the</strong>se and o<strong>the</strong>r kinds of disturbances<br />

affect <strong>for</strong>est nutrient and water cycles,<br />

solar penetration to <strong>the</strong> understory, and plant<br />

and animal food webs. The response of understory<br />

vegetation, fungal-small mammal relationships<br />

(Maser et al. 1978), and owl prey to<br />

various disturbance factors can be positive or<br />

negative, depending on numerous site factors<br />

and <strong>the</strong> successional stage of <strong>the</strong> affected vegetation.<br />

Because <strong>the</strong>se processes are interactive and<br />

affect a number of vegetation attributes, simple<br />

assessments are inadequate. Several vegetation<br />

management tools, including various kinds of<br />

silviculture, risk-abatement <strong>for</strong> fire or insect/<br />

disease damage, prescribed burning, and direct<br />

population control are appropriate in various<br />

combinations.<br />

These disturbance agents should be considered<br />

in developing management strategies <strong>for</strong><br />

owl recovery. Managers must recognize that <strong>the</strong><br />

organisms discussed above and <strong>the</strong>ir effects are<br />

not necessarily or even primarily bad. Certain<br />

natural processes may interfere with short-term<br />

priorities of <strong>for</strong>est management; but <strong>the</strong> perpetuation<br />

of <strong>for</strong>est conditions that support those<br />

priorities may depend on natural processes<br />

continuing in <strong>the</strong> long term. Moreover, conflicting<br />

priorities, or even second- or third-level<br />

priorities may benefit from <strong>the</strong>se organisms.<br />

Evaluations should be based on <strong>the</strong> role <strong>the</strong>se<br />

organisms play in directing succession toward, or<br />

away from, desired future conditions at different<br />

spatiotemporal scales.<br />

Managers, in consultation with specialists,<br />

can use <strong>the</strong>se organisms to strategic advantage in<br />

creating, enhancing, or maintaining habitats <strong>for</strong><br />

owls (and associated biota) in accord with


landscape goals. For example, dwarf mistletoe<br />

creates nest sites <strong>for</strong> owls in Douglas-fir. In some<br />

places, outbreaks of western spruce budworm<br />

eliminate understory host trees, helping to<br />

reduce fuel ladders that carry fires into tree<br />

crowns. These biotic agents of mortality have<br />

thinning effects on tree overstories. Such thinning<br />

affects nutrient and hydrological cycles,<br />

understory vegetation, and availability of prey to<br />

owls.<br />

In summary, we encourage resource managers<br />

to work with <strong>for</strong>est insect and disease specialists<br />

to develop ecological assessments of <strong>the</strong>se<br />

kinds of disturbances at various scales<br />

(Kaufmann et al. 1994). Understanding <strong>the</strong><br />

scientific basis of <strong>for</strong>est change and evolution is<br />

crucial to successful management of <strong>for</strong>est<br />

ecosystems, and <strong>the</strong>re<strong>for</strong>e to recovery of <strong>the</strong><br />

spotted owl.<br />

Degradation Degradation of of of Riparian Riparian Forests<br />

Forests<br />

Riparian <strong>for</strong>ests may also function as important<br />

components of ecosystems supporting<br />

spotted owls. These communities, particularly<br />

mature, multi-layered <strong>for</strong>ests, could be important<br />

linkages between o<strong>the</strong>rwise isolated subpopulations<br />

of spotted owls. They may serve as<br />

direct avenues of movement between mountain<br />

ranges or as stopover sites where drainages bisect<br />

large expanses of landscape that o<strong>the</strong>rwise would<br />

be inhospitable to dispersing owls. Fur<strong>the</strong>r,<br />

historical evidence exists that spotted owls once<br />

nested in such habitats.<br />

Many riparian ecosystems have deteriorated<br />

in <strong>the</strong> Southwest (Cooperrider 1991, Bock et al.<br />

1993, USDI 1994), and <strong>the</strong> loss of riparian<br />

habitat was one of <strong>the</strong> reasons <strong>for</strong> listing <strong>the</strong> owl<br />

(Part I). Dick-Peddie (1993) estimated from<br />

map and air photo data that 96% of <strong>the</strong> Rio<br />

Grande riparian area in New Mexico has been<br />

lost to urbanization, agriculture, water impoundments,<br />

and o<strong>the</strong>r modifications. Gallery <strong>for</strong>ests<br />

that once extended into woodlands, grasslands,<br />

and deserts have significantly declined or deteriorated,<br />

adversely affecting numerous wildlife<br />

populations (Minckley and Clark 1984, Skovlin<br />

1984, Minckley and Rinne 1985, Bock et al.<br />

1993, USDI 1994). Ef<strong>for</strong>ts to improve riparian<br />

and watershed conditions (DeBano and Schmidt<br />

Volume I/Part II<br />

65<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

1989a, 1989b) could facilitate movements of<br />

spotted owls between distant geographic locations<br />

and perhaps even provide nesting habitat.<br />

A wide variety of o<strong>the</strong>r organisms would also<br />

benefit from healthier riparian systems.<br />

TIMBER TIMBER TIMBER HARVEST HARVEST AND<br />

AND<br />

SILVICULTURAL SILVICULTURAL SILVICULTURAL PRACTICES<br />

PRACTICES<br />

Historically, <strong>the</strong> principal objectives of <strong>for</strong>est<br />

management were to derive economic gain and<br />

commodities from <strong>for</strong>ests. Silviculture has great<br />

potential as a tool <strong>for</strong> meeting o<strong>the</strong>r objectives,<br />

however, such as maintaining and developing<br />

<strong>Mexican</strong> spotted owl habitat, alleviating fire risk,<br />

minimizing impacts of insects and disease, and<br />

enhancing various ecological values. In this<br />

section, we review past timber-harvest practices<br />

in <strong>the</strong> Southwest and contrast those practices<br />

with alternatives. Our focus is <strong>the</strong> potential<br />

effects of <strong>the</strong>se practices on <strong>Mexican</strong> spotted<br />

owls.<br />

Historical Historical Perspectives<br />

Perspectives<br />

Past Past Practices<br />

Practices<br />

The primary factors leading to <strong>the</strong> listing of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl were adverse modification<br />

of its habitat as <strong>the</strong> result of even-aged<br />

management and plans to continue this harvest<br />

method as detailed in existing Forest <strong>Plan</strong>s.<br />

Fletcher (1990) reported <strong>the</strong> loss of >325,000 ha<br />

(800,000 acres) of spotted owl habitat within FS<br />

Region 3 as <strong>the</strong> result of human activities,<br />

primarily <strong>for</strong>est management. Silviculture<br />

emphasized even-aged systems which tended to<br />

simplify stand structure and harvest a disproportionate<br />

share of large trees. The Team used past<br />

<strong>for</strong>est inventory data to evaluate <strong>the</strong> change in<br />

<strong>the</strong> size-class distribution of trees from <strong>the</strong> 1960s<br />

to <strong>the</strong> 1980s. The trend that emerged from our<br />

analysis was a substantial increase in <strong>the</strong> density<br />

of trees 12.7-32.8 cm (5-12.9 in) dbh, but a<br />

large decrease in numbers of trees >48.3 cm (19<br />

in) dbh (see below). As discussed by Ganey and<br />

Dick (1995), large trees are an important component<br />

of spotted owl habitat; thus, <strong>the</strong> 20%<br />

decrease in numbers of trees >48.3 cm (19 in)


dbh removed a key habitat component of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. The simplification of stand<br />

structure is not so easily quantified. Given that<br />

mostly even-aged management was used, however,<br />

<strong>the</strong> conclusion of stand simplification is<br />

reasonable.<br />

Forest Forest <strong>Plan</strong>s<br />

<strong>Plan</strong>s<br />

Existing Forest <strong>Plan</strong>s and <strong>the</strong>ir underlying<br />

standards and guidelines are fairly explicit with<br />

respect to <strong>the</strong> silvicultural practices to be used<br />

and <strong>the</strong> expected timber volumes to be extracted.<br />

These Forest <strong>Plan</strong>s articulate classic even-aged<br />

management regimes with regeneration treatments<br />

occurring at 120-year intervals, intermediate<br />

treatments employed to maintain open stand<br />

conditions, and disease-control treatments as<br />

conditions warrant. Thus, management called<br />

<strong>for</strong> fairly frequent entries into a stand. Fur<strong>the</strong>r,<br />

this management system stressed simple stand<br />

structures, decreased residual densities, and<br />

elimination of large, slow-growing, but high<br />

value trees (primarily ponderosa pine and Douglas-fir).<br />

Salvage, sanitation, fuel reductions, and<br />

fuelwood harvest as specified in Forest <strong>Plan</strong>s<br />

combined to reduce numbers of snags, ano<strong>the</strong>r<br />

correlate of spotted owl habitat. In summary,<br />

even-aged management as specified in Forest<br />

<strong>Plan</strong>s is incompatible with maintaining and<br />

developing spotted owl habitat. The Team is<br />

encouraged, however, by recent ef<strong>for</strong>ts by FS<br />

Region 3 to amend <strong>for</strong>est plans to incorporate<br />

<strong>the</strong> recommendations proposed in this <strong>Recovery</strong><br />

<strong>Plan</strong>, and to emphasize uneven-aged management<br />

as <strong>the</strong> preferred silvicultural system in <strong>the</strong><br />

Region.<br />

Habitat Habitat Trends<br />

Trends<br />

Historical and current trends in spotted owl<br />

habitat are presently unknown. Numerous<br />

factors underlie this lack of knowledge, but <strong>the</strong><br />

paucity of reliable vegetation data is <strong>the</strong> most<br />

glaring explanation. This lack of credible data<br />

has not precluded rampant speculation on<br />

habitat trend, however. In general, habitat trend<br />

is perceived in two divergent ways. One view is<br />

that past timber harvest within <strong>the</strong> <strong>for</strong>est types<br />

used by <strong>Mexican</strong> spotted owls has caused a<br />

Volume I/Part II<br />

66<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

dramatic decline in habitat quantity and quality.<br />

Indeed, <strong>the</strong> conclusion of historical habitat loss<br />

coupled with projections <strong>for</strong> additional habitat<br />

loss were <strong>the</strong> primary factors <strong>for</strong> listing <strong>the</strong><br />

subspecies (Part I). The contrary view suggests<br />

that many years of fire exclusion within Southwestern<br />

<strong>for</strong>ests has allowed mixed-conifer <strong>for</strong>est<br />

types to increase at <strong>the</strong> expense of meadows and<br />

fire-disclimax species such as quaking aspen and<br />

ponderosa pine (USDA 1993b, Johnson 1994).<br />

Fur<strong>the</strong>r, Southwestern ponderosa pine <strong>for</strong>ests are<br />

known to be generally denser today than <strong>the</strong>y<br />

were in pre-settlement times (Covington and<br />

Moore 1992, 1994a, b, c). Based on this in<strong>for</strong>mation,<br />

USDA (1993b) concluded that habitat<br />

suitability <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl had<br />

increased. Because of <strong>the</strong>se conflicting views, <strong>the</strong><br />

Team attempted a quantitative evaluation of<br />

habitat trend with respect to <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl.<br />

Data Data Data Availability. Availability.—Limited Availability. sources of data are<br />

available <strong>for</strong> assessing habitat trend. Within<br />

<strong>for</strong>ested types, <strong>for</strong>est inventories from <strong>the</strong> 1960s<br />

(Choate 1966, Spencer 1966) and <strong>the</strong> 1980s<br />

(Conner et al. 1990, Van Hooser et al. 1993)<br />

have been compared by USDA (1993b) and<br />

Johnson (1994) and are used, in part, <strong>for</strong> our<br />

analyses. We admit, however, that differences in<br />

definitions and in how data were collected make<br />

comparisons between <strong>the</strong> 1960s and 1980s data<br />

tenuous, at best (Van Hooser et al. 1993). These<br />

differences include: (1) changes in definitions of<br />

vegetation types; (2) changes in <strong>the</strong> landbase<br />

being sampled, (e.g., changes in wilderness<br />

designation); and (3) changes in sampling<br />

intensity.<br />

The following comparisons are limited to<br />

commercial <strong>for</strong>est lands within <strong>the</strong> States of<br />

Arizona and New Mexico on a per hectare basis.<br />

Thus, all <strong>for</strong>est types are included but, unlike<br />

USDA (1993b) and Johnson (1994) we do not<br />

extrapolate <strong>the</strong> data to unsampled <strong>for</strong>ested lands<br />

such as wilderness areas. There<strong>for</strong>e, our analyses<br />

focus on changes on commercial <strong>for</strong>est lands<br />

where data exist. Because of differences in land<br />

designations (i.e., commercial timber land<br />

becoming wilderness between <strong>the</strong> two sampling<br />

periods), comparisons of raw values are potentially<br />

misleading. Thus, our comparisons are


primarily restricted to evaluations of proportions.<br />

To compare stand structure, we used<br />

relative frequencies of trees by size class. We<br />

reiterate that caution is warranted when inferring<br />

conclusions from <strong>the</strong>se data, but submit that<br />

some gross generalizations are possible.<br />

Trends Trends in in Forest Forest Landbase Landbase and and Timber<br />

Timber<br />

Volume.— Volume.—Total Volume.— <strong>for</strong>ested land increased from<br />

4,516,000 to 4,750,000 ha (11,160,000 to<br />

11,738,000 acres) from <strong>the</strong> 1960s to <strong>the</strong> 1980s,<br />

roughly a 5% increase. The commercial <strong>for</strong>est<br />

landbase decreased by approximately 15%<br />

(624,000 ha [1,541,000 acres]), however, and<br />

reserved <strong>for</strong>ested lands increased by 858,000 ha<br />

(2,119,000 acres). Growing stock (i.e., <strong>the</strong><br />

harvestable volume) on commercial lands decreased<br />

from 12,707 MMCF to 11,549 MMCF.<br />

This decrease is not surprising given <strong>the</strong> volume<br />

of timber harvested on commercial <strong>for</strong>est lands<br />

and <strong>the</strong> decrease in <strong>the</strong> amount of commercial<br />

<strong>for</strong>est lands from <strong>the</strong> 1960s to <strong>the</strong> 1980s.<br />

Trends Trends in in in Forest Forest Types. Types.—Within Types. <strong>the</strong> commercial<br />

landbase, mixed-conifer <strong>for</strong>ests comprised<br />

approximately 11% of total area in <strong>the</strong> 1960s<br />

Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able II.D.1. II.D.1. Changes in <strong>the</strong> area (ha X 1,000 [acres x 1,000]) and distribution of <strong>for</strong>est types from<br />

<strong>the</strong> 1960s to 1980s on commercial <strong>for</strong>est lands within Arizona and New Mexico. Data from Choate<br />

(1966), Spencer (1966), Conner et al. (1990), Van Hooser et al. (1993).<br />

__________________________________________________________________________________________<br />

__________________________________________________________________________________________<br />

Landbase Landbase Propor opor oportion opor tion of of Landbase Landbase Propor opor oportion opor tion of of Change Change in<br />

in<br />

For or orest orest<br />

est Type ype ype in in 1960s 1960sa<br />

1960s 1960s 1960s Landbase Landbaseb<br />

in in 1980s 1980s 1980sa<br />

1980s 1980s Landbase Landbaseb<br />

P PPropor<br />

P opor oportion opor tion c<br />

__________________________________________________________________________________________<br />

__________________________________________________________________________________________<br />

Ponderosa Pine 3,234 78 2,530 72 -6<br />

[7,992] [6,252]<br />

Mixed-conifer 475 11 709 20 9<br />

[1,173] [1,752]<br />

Spruce-fir 257 6 201 6 0<br />

[635] [496]<br />

Quaking aspen 180 4 81 2 -2<br />

[446] [201]<br />

Total otal 4,146 4,146 100 100 3,523 3,523<br />

100<br />

100<br />

[10,246] [10,246] [10,246]<br />

[8,701] [8,701]<br />

[8,701]<br />

__________________________________________________________________________________________<br />

a Landbase in hectares [acres] covered by <strong>the</strong> <strong>for</strong>est type.<br />

b Proportion of <strong>the</strong> total <strong>for</strong>ested landbase belonging to <strong>the</strong> <strong>for</strong>est type.<br />

c (Prop. 1960s landscape)-(Prop. 1980s landscape).<br />

67<br />

and 20% of <strong>the</strong> total area in <strong>the</strong> 1980s, a 9%<br />

increase (Table II.D.1). Possible explanations <strong>for</strong><br />

this change include: (1) increasing invasion of<br />

mixed-conifer species (presumably Douglas-fir<br />

and white fir) into o<strong>the</strong>r types, such as meadows;<br />

(2) more liberal definitions of mixedconifer<br />

(i.e., includes types previously classified<br />

as something else); (3) quaking aspen giving way<br />

to o<strong>the</strong>r species in <strong>the</strong> absence of fire; and (4)<br />

selective harvest of ponderosa pine, leaving<br />

residual <strong>for</strong>ests composed primarily of o<strong>the</strong>r<br />

conifer species. Any of <strong>the</strong>se reasons may explain<br />

<strong>the</strong> perceived changes of <strong>for</strong>est type; probably all<br />

of <strong>the</strong>se and o<strong>the</strong>r factors contributed to some<br />

degree. We speculate that classification changes<br />

account <strong>for</strong> most of <strong>the</strong> change, and that selective<br />

removal of ponderosa pine and <strong>the</strong> succession<br />

of quaking aspen stands to mixed-conifer<br />

are also plausible short-term explanations.<br />

Conversely, we have difficulty accepting that<br />

encroachment of mixed-conifer species into<br />

o<strong>the</strong>r <strong>for</strong>ested types was responsible <strong>for</strong> more<br />

than a relatively small portion of this change<br />

within <strong>the</strong> twenty-year period. Thus, any generalizations<br />

concerning changes in <strong>for</strong>est types and<br />

any actions proposed to reverse <strong>the</strong>se trends


Volume I/Part II<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able II.D.2. II.D.2. Changes in <strong>the</strong> density (trees/ha [trees/acre]) and distribution of tree size classes from<br />

<strong>the</strong> 1960s to 1980s on commercial <strong>for</strong>est lands within Arizona and New Mexico. Data from Choate<br />

(1966), Spencer (1966), Conner et al. (1990), Van Hooser et al. (1993).<br />

Tree ee Density ensity Propor opor oportion opor tion of of Density ensity Propor opor oportion opor tion of of Change Change in in Density ensity<br />

Siz iz ize iz e Class Class Class in in in 1960s 1960s 1960sa<br />

1960s 1960s Total otal b in in 1980s 1980sa<br />

1980 1980 Total otal b Propor opor oportion opor tion c Change<br />

Change<br />

Change d<br />

2.5-12.5 cm 146.2 62.5 134.2 53.7 -8.8 -8.3<br />

[1.0-4.9 in] [59.2] [54.3]<br />

12.6-32.8 cm 70.3 30.0 98.5 39.4 9.4 40.2<br />

[5.0-12.9 in] [28.5] [39.9]<br />

32.9-48.0 cm 12.1 5.2 13.0 5.2 0.0 7.5<br />

[13.0-18.9 in] [4.9] [5.3]<br />

>48 cm 5.4 2.3 4.3 1.7 -0.6 -20.4<br />

[>19 in] [2.2] [1.7]<br />

a Tree density in no./ha [no./acre]<br />

b Proportion of <strong>the</strong> total number of trees within that size class.<br />

c (Prop. 1960s total)-(Prop. 1980s total).<br />

d<br />

([Prop. 1960s total]-[Prop. 1980s total])/(Prop. 1960s total).<br />

must acknowledge this uncertainty, and should<br />

consider all plausible explanations <strong>for</strong> <strong>the</strong>se<br />

trends.<br />

Trends Trends in in Size-class Size-class Size-class Distributions.<br />

Distributions.—We Distributions. noted<br />

a change in <strong>the</strong> size-class distribution of trees on<br />

commercial <strong>for</strong>est lands of Arizona and New<br />

Mexico (Table II.D.2). Sapling-sized trees (2.5-<br />

12.5 cm [1-4.9 in] dbh) decreased in both<br />

absolute density and in relative contribution to<br />

<strong>the</strong> size-class distribution; trees 12.6-31 cm<br />

(5-12 in) dbh increased in density by 40% and<br />

in relative proportion of <strong>the</strong> size class distribution<br />

by >9%; and trees in <strong>the</strong> 31-48 cm (13-19<br />

in) size class increased in density but not in<br />

relative proportion of <strong>the</strong> tree distribution.<br />

Finally, <strong>the</strong> density of large trees (>48 cm [19 in]<br />

dbh) decreased from 2.3 to 1.7 trees/ha (0.9 to<br />

0.7 trees/ac), a 20% decline. This decrease in<br />

large trees would be expected given past timber<br />

harvest practices which emphasized harvest of<br />

<strong>the</strong> large trees. Possible explanations <strong>for</strong> <strong>the</strong><br />

increase in smaller stems include <strong>the</strong> growth of<br />

regeneration, limited pre-commercial thinning,<br />

fire suppression, and <strong>the</strong> lack of interest by <strong>the</strong><br />

<strong>for</strong>est industry in <strong>the</strong> smaller-sized stems.<br />

Summary Summary Summary of of of Recent Recent Recent Habitat Habitat Habitat Trends. Trends. Trends.—Our<br />

Trends. Trends.<br />

analyses indicate that between <strong>the</strong> 1960s and<br />

68<br />

1980s (1) total <strong>for</strong>ested acres increased, (2)<br />

mixed-conifer types apparently covered more of<br />

<strong>the</strong> landbase (but see above cautions regarding<br />

this conclusion), and (3) densities of large trees<br />

declined. Although <strong>the</strong> amount of total <strong>for</strong>ested<br />

land has increased and <strong>the</strong> amount of mixedconifer<br />

<strong>for</strong>est may have increased, we doubt that<br />

<strong>the</strong> amount of <strong>Mexican</strong> spotted owl habitat has<br />

increased concomitantly. Given <strong>the</strong> 20-yr period<br />

between inventories, most of <strong>the</strong>se additional<br />

acres are likely in early successional stages and<br />

unlikely to possess <strong>the</strong> habitat characteristics<br />

used by spotted owls. Conversely, <strong>the</strong> 20%<br />

decrease in <strong>the</strong> density of large trees is an alarming<br />

negative trend with respect to a very critical<br />

component of spotted owl habitat.<br />

Silvicultural Silvicultural Practices<br />

Practices<br />

and and Forest Forest Forest Management<br />

Management<br />

Four common <strong>for</strong>est structures occur naturally<br />

or by silvicultural ef<strong>for</strong>ts. These structures<br />

are even-aged, balanced uneven-aged, irregular<br />

uneven-aged, and even-aged/uneven-aged<br />

stratified mixtures. Even-aged stands, <strong>for</strong> <strong>the</strong><br />

most part, are characterized by most trees being<br />

approximately <strong>the</strong> same age. The general convention<br />

is that <strong>the</strong> spread of ages within <strong>the</strong><br />

stand are within approximately 20% of <strong>the</strong>


specified rotation age. Two types of uneven-aged<br />

stands are balanced and irregular stands. In both<br />

cases, at least three distinct age classes exist. A<br />

balanced uneven-aged stand equates to each age<br />

class occupying roughly equal areas. Distribution<br />

by diameter class approximates a reverse, jshaped<br />

curve. Irregular uneven-aged stands have<br />

some age and associated diameter classes missing<br />

across <strong>the</strong> possible range of ages and diameters.<br />

A two-storied stand, one with two distinct ageclass<br />

and diameter distributions, is nei<strong>the</strong>r evenor<br />

uneven-aged, but is intermediate between <strong>the</strong><br />

two. Stratified mixtures occur where trees are<br />

essentially even-aged, but differences in growth<br />

rates and shade tolerance among tree species<br />

result in multiple canopy strata. This structure<br />

also occurs when selective regeneration of shadetolerant<br />

species or high site productivity leads to<br />

heterogeneous age and diameter distributions.<br />

In much of <strong>the</strong> mixed-conifer type in <strong>the</strong> Southwest,<br />

<strong>the</strong> stratified-mixture of stand structure<br />

appears to be relevant to habitats used by spotted<br />

owls.<br />

Silviculture<br />

Silviculture<br />

Silviculture has been variously defined as (1)<br />

<strong>the</strong> art of producing and tending a <strong>for</strong>est, (2) <strong>the</strong><br />

application of knowledge of silvics in <strong>the</strong> treatment<br />

of a <strong>for</strong>est, and (3) <strong>the</strong> <strong>the</strong>ory and practice<br />

of controlling <strong>for</strong>est establishment, composition,<br />

structure and growth (Smith 1986). In a general<br />

sense, silviculture is <strong>the</strong> practice of managing<br />

<strong>for</strong>est establishment, composition, structure, and<br />

growth to meet stated objectives. Thus, silviculture<br />

should be regarded as a system of treatments<br />

and not solely <strong>the</strong> practice of removing trees<br />

from a stand.<br />

In <strong>the</strong> Southwest, two broad classifications<br />

of silvicultural systems, based on methods of<br />

reproduction and resulting age-class mixes of<br />

<strong>for</strong>ested stands, are even-aged and uneven-aged<br />

management. These are reviewed below; again,<br />

our focus is <strong>the</strong> potential <strong>the</strong>se systems have <strong>for</strong><br />

developing spotted owl habitat.<br />

Even-aged Even-aged Management.<br />

Management.—Even-aged Management.<br />

management<br />

has been used commonly in Southwestern<br />

<strong>for</strong>ests. Reasons <strong>for</strong> its popularity are based both<br />

on ecology and economics. Ecologically, even-<br />

Volume I/Part II<br />

69<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

aged systems favor species with limited shade<br />

tolerance, such as quaking aspen, oaks, and<br />

lodgepole pine. Shade tolerance is <strong>the</strong> ability to<br />

reproduce and grow under <strong>the</strong> shade of larger,<br />

taller trees. Shade-tolerant species typically<br />

include true firs in Southwestern <strong>for</strong>ests. Ponderosa<br />

pine is considered to be of intermediate<br />

shade tolerance, but tending toward <strong>the</strong> intolerant<br />

side. Economically, even-aged management<br />

is more efficient when considering short-term<br />

costs of site utilization, sale preparation, transportation<br />

systems, harvesting, and slash reduction.<br />

Fur<strong>the</strong>r, even-aged systems are easier to<br />

model, administer, and track over time.<br />

Regeneration methods within even-aged<br />

systems of <strong>the</strong> Southwest include shelterwood,<br />

clearcutting, and seed tree methods. The<br />

shelterwood method typically has a series of<br />

cuttings. The first treatment in mature stands is<br />

to stimulate cone and seed production <strong>for</strong><br />

regeneration. This is followed by a series of<br />

treatments that remove <strong>the</strong> larger, older stems as<br />

regeneration matures. Variations on <strong>the</strong> general<br />

method include irregular shelterwood and<br />

group-shelterwood. Clearcutting involves <strong>the</strong><br />

removal of <strong>the</strong> entire stand in one cutting.<br />

Reproduction is obtained artificially by seeding<br />

or planting, or naturally by seeding from adjacent<br />

stands. This method is appropriate <strong>for</strong><br />

shade-intolerant species. In appearance,<br />

clearcutting is indistinguishable from <strong>the</strong> coppice-<strong>for</strong>est<br />

method of regeneration <strong>for</strong> quaking<br />

aspen, where reproduction is obtained from<br />

suckering of sub-terrain clones. The seed-tree<br />

method resembles clearcutting except that a few<br />

trees are left to provide a source of seed within<br />

<strong>the</strong> treated area. Of <strong>the</strong>se three, <strong>the</strong> shelterwood<br />

method is used most commonly in <strong>the</strong> Southwest;<br />

clearcut and seed-tree methods are used<br />

infrequently.<br />

Variations of even-aged management are<br />

used throughout <strong>the</strong> Southwest, but all share <strong>the</strong><br />

following characteristics. A predetermined time<br />

<strong>for</strong> regeneration of <strong>the</strong> stand is set a priori; this<br />

regeneration time can vary. The FS Region 3<br />

generally schedules regeneration treatments from<br />

100 to 120 years of stand age, an age when trees<br />

are expected to reach 45.7 cm (18 in) dbh as <strong>the</strong><br />

maximum size. The management objective is to<br />

maximize total volume over time while provid-


ing a “saw-timber” sized product. Residual stand<br />

density can be controlled by thinning at <strong>the</strong>oretically<br />

scheduled entries in <strong>the</strong> stand. This<br />

enables <strong>the</strong> capture of mortality on a semiregular<br />

basis and provides intermediate revenues<br />

from timber harvest.<br />

Even-aged stand structures are not used to<br />

any great extent by <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Fur<strong>the</strong>r, with its intent to promote uni<strong>for</strong>mity in<br />

tree age, size, spacing, and density, even-aged<br />

management generally would not be a preferred<br />

system <strong>for</strong> short-term development of spotted<br />

owl habitat. Even-aged management may be<br />

appropriate to maintain quaking aspen within<br />

<strong>the</strong> mixed-conifer type, however. Quaking aspen<br />

is typically even-aged in its earlier stages of stand<br />

development. Because of its extreme shade<br />

intolerance and requirements <strong>for</strong> elevated soil<br />

temperatures <strong>for</strong> sprouting, quaking aspen<br />

should be managed under even-aged systems.<br />

Later seral stages of quaking aspen that include a<br />

mixed-conifer understory appear both as a<br />

simple mixture of an even-aged overstory and<br />

uneven-aged understory, or as a stratified mixture.<br />

As decadent quaking aspen stands are<br />

replaced by <strong>the</strong> shade-tolerant conifers, spotted<br />

owl nesting/roosting habitat begins to develop.<br />

In summary, even-aged management has limited<br />

potential with respect to developing <strong>the</strong> types of<br />

stand structures used by spotted owls. Never<strong>the</strong>less,<br />

even-aged management is a critical tool to<br />

meet specific objectives, and can meet certain<br />

ecosystem objectives if employed at <strong>the</strong> proper<br />

scale. Perhaps <strong>the</strong> primary objection to its use in<br />

<strong>the</strong> past was its uni<strong>for</strong>m, widescale application<br />

across <strong>the</strong> Southwest.<br />

Uneven-aged Uneven-aged Management.<br />

Management.—Uneven-aged<br />

Management.<br />

management entails <strong>the</strong> removal of timber in all<br />

size classes on a periodic basis so that regeneration<br />

is continuously established over time, and<br />

stand size-class distribution is regulated. Uneven-aged<br />

management is loosely based on <strong>the</strong><br />

premise that density control across a range of<br />

diameter classes will ensure growth of stems over<br />

time to a set maximum diameter, while ensuring<br />

regeneration of species at regular intervals over<br />

time. Only one reproduction method, <strong>the</strong><br />

selection method, is used with uneven-aged<br />

management. The selection method provides<br />

Volume I/Part II<br />

70<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

openings in <strong>the</strong> stand to enable regeneration to<br />

occur. Simultaneous with <strong>the</strong> selection and<br />

harvest of trees to provide growing space <strong>for</strong><br />

regeneration, trees across all diameter classes are<br />

thinned to ensure <strong>the</strong> desired distribution of<br />

size- and age-classes within <strong>the</strong> stand.<br />

Two variations of <strong>the</strong> selection method are<br />

individual tree selection and group selection.<br />

Individual tree selection, as <strong>the</strong> name implies,<br />

involves <strong>the</strong> removal of single, scattered trees.<br />

This method generally favors shade-tolerant<br />

species, but this is also a function of <strong>the</strong> residual<br />

stocking levels. Group selection entails <strong>the</strong><br />

removal of a small patch of trees; <strong>the</strong> width of<br />

<strong>the</strong> patch is usually less than twice <strong>the</strong> height of<br />

<strong>the</strong> dominant (i.e., largest) tree. This is somewhat<br />

analogous to a very small clearcut, but <strong>the</strong><br />

difference between <strong>the</strong> group selection and <strong>the</strong><br />

clearcut method is in <strong>the</strong> spatial scale of application.<br />

Group selection is used to create a balance<br />

of age- or size-classes in small contiguous groups<br />

resulting in a mosaic within a stand. In contrast,<br />

even-aged methods are typically applied to an<br />

entire stand. Group selection can be used to<br />

promote <strong>the</strong> establishment and growth of shadeintolerant<br />

trees since <strong>the</strong>re is opportunity to<br />

reduce localized residual densities and <strong>the</strong><br />

amount of area shaded.<br />

Group selection offers a number of advantages<br />

<strong>for</strong> <strong>the</strong> development of potential spotted<br />

habitat over single-tree selection techniques.<br />

Application of <strong>the</strong> group selection method could<br />

provide a mosaic of many small even-aged or<br />

two-storied groups across a <strong>for</strong>est stand. Regeneration<br />

of shade-intolerant species is possible<br />

where a reproduction source, ei<strong>the</strong>r clones or<br />

seeds, is present. With respect to insect and<br />

disease problems, management options increase<br />

<strong>for</strong> both suppression and prevention, especially<br />

in mixed-species stands. Edge effects found at<br />

group interfaces can provide structural features<br />

and openings that mimic gap-phase regeneration,<br />

and provide early-seral vegetation <strong>for</strong> prey<br />

species (Ward and Block 1995). In some cases,<br />

group-selection methods may result in less<br />

residual damage to <strong>the</strong> stand as <strong>the</strong> result of<br />

logging activities than single-tree selection.<br />

Uneven-aged silvicultural practices predominate<br />

on Tribal lands where commercial timber<br />

harvest exists. Reasons <strong>for</strong> <strong>the</strong> emphasis on this


silvicultural system include aes<strong>the</strong>tics, providing<br />

<strong>for</strong>est cover over all lands simultaneously, <strong>the</strong><br />

perception that it provides a more even-flow of<br />

products than even-aged management, and <strong>the</strong><br />

fact that it allows continuous regeneration.<br />

We have not been able to assess <strong>the</strong> effects of<br />

classic uneven-aged management on <strong>Mexican</strong><br />

spotted owl habitat because we were unable to<br />

acquire data <strong>for</strong> most areas where uneven-aged<br />

management is practiced on a large scale. However,<br />

based upon our understanding of <strong>the</strong><br />

application of uneven-aged systems, stand<br />

density is often kept at a fairly low level, seldom<br />

exceeding 18 m 2 /ha (80 ft 2 /acre) of basal area.<br />

These low residual stand densities allow <strong>for</strong><br />

regeneration and growth of ponderosa pine.<br />

Uneven-aged systems, whe<strong>the</strong>r <strong>the</strong>y retain<br />

individual trees or groups of trees, allow <strong>for</strong> <strong>the</strong><br />

development of multiple canopy levels, a key<br />

component of <strong>Mexican</strong> spotted owl habitat.<br />

However, Ganey and Dick (1995) demonstrate<br />

clearly that owl habitat typically also includes<br />

significant numbers of large trees. These large<br />

trees may not be retained where uneven-aged<br />

management is applied in this fashion.<br />

In summary, uneven-aged management has<br />

some promise <strong>for</strong> providing stands exhibiting<br />

characteristics of spotted owl habitat. As currently<br />

practiced, however, uneven-aged management<br />

results in large acreages of low-density<br />

stands, numerous road openings, and <strong>the</strong> eventual<br />

eradication of large diameter stems. Although<br />

nei<strong>the</strong>r <strong>the</strong> short- or <strong>the</strong> long-term<br />

effects of <strong>the</strong>se applications on spotted owls are<br />

known, this type of application may not be <strong>the</strong><br />

best option <strong>for</strong> producing spotted owl habitat.<br />

Development Development or or or maintenance maintenance of of stratified<br />

stratified<br />

mixtures. mixtures. mixtures.—Stratified mixtures. mixtures can originate in<br />

various ways, including stand-replacing events<br />

(e.g., crown fire, even-aged management) that<br />

may or may not leave remnant stems. The<br />

establishment of stratified mixtures is not likely<br />

in mixed-conifer types within <strong>the</strong> short time<br />

frame (10-15 years) of this <strong>Recovery</strong> <strong>Plan</strong>.<br />

Maintenance of such stands could be considered<br />

in <strong>the</strong> short-term, however, and development of<br />

stratified mixtures could be incorporated in<br />

longer-term management plans.<br />

Volume I/Part II<br />

71<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Management <strong>for</strong> stratified mixtures must<br />

consider <strong>the</strong> mix of species along <strong>the</strong> continuum<br />

of shade tolerance. Thus, any regeneration<br />

ef<strong>for</strong>ts should be designed to have enough<br />

openings of sufficient size <strong>for</strong> seedling/sprouting<br />

establishment and release. Openings can be<br />

accomplished by group selection cuts favoring<br />

retention of shade-intolerant species, or selection<br />

of individual trees adjacent to stems that could<br />

provide a seed source <strong>for</strong> regeneration. One<br />

technique to consider is small-scale seed tree cuts<br />

(perhaps to be thought of as group seed-tree<br />

selection cuts), which would provide both a seed<br />

source and trees <strong>for</strong> ultimate snag development.<br />

This method could maintain shade-intolerant<br />

species, but would not be intrusive enough to<br />

produce even-aged structure throughout <strong>the</strong><br />

stand. Within stratified mixtures, intermediate<br />

treatments including pre-commercial and<br />

commercial thinning could be beneficial in<br />

increasing <strong>the</strong> growth of residual trees. At some<br />

point, however, fur<strong>the</strong>r treatments should be<br />

deferred, and natural stand maturation and<br />

succession should be allowed to proceed until<br />

ei<strong>the</strong>r (1) <strong>the</strong> stand is no longer spotted owl<br />

habitat; or (2) <strong>the</strong> stand can be replaced by<br />

habitat (preferably occupied by spotted owls)<br />

that has been developed elsewhere.<br />

Conclusions<br />

Conclusions<br />

Clearly, recent <strong>for</strong>est management practices<br />

and those detailed in existing Forest <strong>Plan</strong>s are<br />

not beneficial to <strong>Mexican</strong> spotted owls. Reliance<br />

on traditional <strong>for</strong>est management and silvicultural<br />

techniques may no longer be possible, not<br />

only with respect to <strong>the</strong> conservation of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl but also with respect to<br />

maintaining o<strong>the</strong>r ecosystem attributes. New<br />

approaches must be developed that ensure <strong>the</strong><br />

long-term provision of owl habitat and <strong>the</strong><br />

maintenance of ecosystem structure and function.<br />

Traditional approaches will still have <strong>the</strong>ir<br />

role, but perhaps used in slightly different ways<br />

and with different intensities. Innovative applications<br />

of uneven-aged management may be<br />

particularly useful in developing and maintaining<br />

spotted owl habitat. In addition, particular<br />

applications of uneven-aged management may<br />

be useful in maintaining habitat conditions <strong>for</strong>


<strong>the</strong> owl where <strong>the</strong>y exist. In some cases, <strong>the</strong><br />

application of even-aged management systems<br />

may also be appropriate, so future <strong>for</strong>est management<br />

should not preclude <strong>the</strong> use of evenaged<br />

management.<br />

Volume I/Part II<br />

GRAZING<br />

GRAZING<br />

Grazing by livestock and wildlife (e.g., elk,<br />

deer) occurs throughout <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Depending on <strong>the</strong> intensity,<br />

grazing has <strong>the</strong> potential to influence habitat<br />

composition and structure, and affect food<br />

availability and diversity <strong>for</strong> <strong>the</strong> owl. However,<br />

predicting <strong>the</strong> magnitude of grazing effects on<br />

spotted owls and <strong>the</strong>ir habitat, and evaluating<br />

management options requires a better understanding<br />

of <strong>the</strong> relationship between spotted owl<br />

habitat and grazing.<br />

Specific studies that document <strong>the</strong> effects of<br />

livestock and wildlife grazing on spotted owl<br />

habitat have not been conducted. Until specific<br />

in<strong>for</strong>mation is available, <strong>the</strong> potential effects on<br />

<strong>the</strong> owl of grazing and trampling of vegetation<br />

must be identified and considered to <strong>the</strong> extent<br />

possible. For example, livestock and wildlife may<br />

not impact spotted owl roost and nest sites<br />

immediately, but could alter riparian habitats by<br />

reducing, eliminating, or suppressing regeneration.<br />

In time, reduced regeneration could limit<br />

<strong>the</strong> development of overstory structure needed<br />

<strong>for</strong> nesting, roosting, and o<strong>the</strong>r life history<br />

needs, as well as jeopardize <strong>the</strong> sustainability of<br />

<strong>the</strong>se habitat types.<br />

Grazing can alter a plant community directly,<br />

indirectly, or both. Direct alterations may<br />

be as obvious as plant removal by consumption<br />

or as subtle as removal by trampling. Indirect<br />

alterations may be as straight<strong>for</strong>ward as loss of<br />

seed source or as insidious as damaged soil<br />

(Dwyer et al. 1984, Kauffman and Krueger<br />

1984, Fleischner 1994). Moderate to heavy<br />

grazing can reduce plant density, cover, biomass,<br />

vigor, and regeneration ability. Collectively, <strong>the</strong>se<br />

factors can alter <strong>the</strong> relative composition and<br />

structure of grass, <strong>for</strong>b, shrub, and tree components<br />

in an area (Hanley and Page 1982,<br />

Zimmerman and Neuenschwander 1984, Schulz<br />

and Leininger 1990, Milchunas and Lauenroth<br />

1993). Within conifer <strong>for</strong>ests, grazing can<br />

72<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

remove or greatly reduce grasses and <strong>for</strong>bs,<br />

<strong>the</strong>reby allowing large numbers of conifer<br />

seedlings to become established because of<br />

reduced competition <strong>for</strong> water and nutrients and<br />

reduced allelopathy. Establishment of large<br />

numbers of seedlings coupled with <strong>the</strong> reduction<br />

in light ground fuels (i.e., grasses and <strong>for</strong>bs) may<br />

act synergistically with fire suppression to<br />

contribute to dense overstocking of ladder fuels.<br />

This dense overstocking can alter <strong>for</strong>est structure<br />

and composition and degrade spotted owl and<br />

prey habitats while increasing risks of standreplacing<br />

fires.<br />

Beyond <strong>the</strong> effects of grazing on plants,<br />

livestock activity can increase duff layers, accelerate<br />

decomposition of woody material, produce<br />

compacted soils, damage stream banks and<br />

channels, and damage lake shores (Kennedy<br />

1977, Blackburn 1984, Kauffman and Krueger<br />

1984, Skovlin 1984, Clary and Webster 1989).<br />

The combination of <strong>the</strong>se changes to <strong>the</strong> biotic<br />

and physical landscapes also affects plant community<br />

composition, structure, and vigor. If<br />

such changes occur in or near areas used by<br />

spotted owls, <strong>the</strong>n grazing can influence <strong>the</strong> owl.<br />

Those influences can be manifested by altering<br />

(1) prey availability, (2) susceptibility of spotted<br />

owl habitat to fire, (3) <strong>the</strong> health and condition<br />

of riparian communities; and (4) development of<br />

habitat. We summarize below <strong>the</strong> major suspected<br />

influences of grazing on <strong>Mexican</strong> spotted<br />

owls.<br />

1. For <strong>the</strong> <strong>Mexican</strong> spotted owl, prey<br />

availability is determined by <strong>the</strong> distribution,<br />

abundance, and diversity of prey<br />

and by <strong>the</strong> owl’s ability to capture it.<br />

Grazing may influence prey availability<br />

in dissimilar ways. For example, grazing<br />

that reduces dense grass cover can create<br />

favorable habitat conditions <strong>for</strong> deer<br />

mice while creating unfavorable conditions<br />

<strong>for</strong> voles, meadow jumping mice,<br />

and shrews (Medin and Clary 1990,<br />

Schultz and Leininger 1991). This<br />

change might decrease prey diversity<br />

(Medin and Clary 1990, Hobbs and<br />

Huenneke 1992). A diverse prey base can<br />

provide a more predictable food resource<br />

<strong>for</strong> <strong>the</strong> owls over time, because popula-


Volume I/Part II<br />

tions of many small mammals fluctuate<br />

asynchronously. Conversely, short-term<br />

removal of grass and shrub cover may<br />

improve conditions <strong>for</strong> <strong>the</strong> owl to detect<br />

and capture prey. Long-term loss of<br />

grasses, <strong>for</strong>bs, and shrubs may promote<br />

tree growth and cover that could decrease<br />

prey abundance. Thus, grazing can pose<br />

ecological tradeoffs.<br />

2. Grazing that significantly reduces herbaceous<br />

ground cover and increases shrubs<br />

and small trees can decrease <strong>the</strong> potential<br />

<strong>for</strong> beneficial low-intensity ground fires<br />

while increasing <strong>the</strong> potential <strong>for</strong><br />

destructive high-intensity vertical fires<br />

(Zimmerman and Neuenschwander<br />

1984). Low-intensity ground fires<br />

prevent fuel accumulation, stimulate<br />

nutrient cycling, promote grasses and<br />

<strong>for</strong>bs, discourage shrubs and trees, and<br />

perpetuate <strong>the</strong> patchiness that supports<br />

small mammal diversity. Catastrophic<br />

fire reduces or eliminates <strong>for</strong>aging,<br />

wintering, dispersal, roosting, and<br />

nesting habitat components.<br />

3. Excessive grazing in riparian areas can<br />

reduce or eliminate important shrub,<br />

tree, <strong>for</strong>b, and grass cover, all of which in<br />

some capacity support <strong>the</strong> owl or its<br />

prey. Excessive grazing can also physically<br />

damage stream channels and banks<br />

(Ames 1977, Kennedy 1977, Kauffman<br />

et al. 1983, Blackburn 1984, Slovkin<br />

1984, Clary and Webster 1989, Platts<br />

1990.) Deterioration of riparian vegetation<br />

structure can allow channel widening.<br />

This event, in turn, elevates water<br />

and soil temperatures and thus evaporation<br />

and lowering of water tables, plus it<br />

significantly increases <strong>the</strong> potential <strong>for</strong><br />

accelerated flood damage (Platts 1990).<br />

These processes alter <strong>the</strong> microclimate<br />

and vegetative development of riparian<br />

areas, potentially impairing its use by<br />

spotted owls.<br />

4. Excessive grazing, sustained <strong>for</strong> long<br />

periods, can inhibit or retard an area’s<br />

73<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

ability to produce or eventually mature<br />

into habitat <strong>for</strong> <strong>the</strong> owl or its prey. This<br />

will probably prove to be an inevitable<br />

consequence of <strong>the</strong> events and processes<br />

described above.<br />

The potential <strong>for</strong> grazing to influence<br />

various components of spotted owl habitat<br />

cannot be ignored. However, current predictions<br />

of grazing effects on plant communities as <strong>the</strong>y<br />

relate to <strong>the</strong> owl are inexact. Thus, <strong>the</strong> integration<br />

of spotted owl needs and grazing management<br />

will require coordination, and an interactive<br />

and adaptive approach between protection,<br />

restoration, and management.<br />

RECREATION<br />

RECREATION<br />

Recreational activities may affect <strong>Mexican</strong><br />

spotted owls directly by disturbing nests, roosts,<br />

or <strong>for</strong>aging sites. Disturbance may occur indirectly<br />

through altered habitat caused by trampling<br />

of vegetation, soil damage, or both. Developing<br />

new recreation facilities or expanding<br />

existing facilities, such as campgrounds and<br />

trails, may alter spotted owl habitat and habitat<br />

use and perpetuate disturbance impacts caused<br />

by recreation.<br />

If a given recreational activity does not cause<br />

habitat alteration, <strong>the</strong> Team assumes that that<br />

activity generally has relatively low impact<br />

potential with respect to spotted owls. However,<br />

exceptions may exist in local situations or certain<br />

RUs where <strong>the</strong> level of recreational activities is<br />

high. Essentially, <strong>the</strong> determining factor of an<br />

activity’s impact on spotted owls is a combination<br />

of its location, intensity, frequency, and<br />

duration ra<strong>the</strong>r than simply its character.<br />

Types Types of of Recreation<br />

Recreation<br />

Recreational activities fall into several categories;<br />

<strong>the</strong> number, size, and intensity of such<br />

activities will vary with location. The following<br />

general categories include most widespread<br />

recreational activities that might affect spotted<br />

owls and <strong>the</strong>ir habitat.


Camping Camping<br />

Camping<br />

Although <strong>the</strong> effects of camping on spotted<br />

owls have not been studied, disruption of nesting,<br />

roosting, and <strong>for</strong>aging activities is a distinct<br />

possibility. The character of camping varies<br />

dramatically, however. One person may camp<br />

alone in a small tent, whereas o<strong>the</strong>rs may camp<br />

in groups with motorhomes. The disparity in<br />

character does not necessarily translate to disturbance<br />

potential, however. One person camping<br />

in a nest grove could be more disruptive than 12<br />

people camping in a <strong>for</strong>aging area. There<strong>for</strong>e,<br />

blanket generalizations about <strong>the</strong> impacts of<br />

camping activities are inappropriate. These<br />

activities should be assessed on a case-by-case<br />

basis, considering factors such as <strong>the</strong> location of<br />

<strong>the</strong> activity relative to <strong>the</strong> owls, <strong>the</strong> number of<br />

individuals involved, <strong>the</strong> type of group involved,<br />

and <strong>the</strong> frequency and duration of <strong>the</strong> activity.<br />

Hiking<br />

Hiking<br />

Hiking is typically a short-term activity, and<br />

may bring a person into and out of an owl’s<br />

presence relatively quickly. Most spotted owls<br />

appear to be relatively undisturbed by small<br />

groups (


Volume I/Part II<br />

SUMMARY<br />

SUMMARY<br />

Part III of this <strong>Recovery</strong> <strong>Plan</strong> outlines<br />

management guidelines to alleviate threats to <strong>the</strong><br />

spotted owl. These recommendations are based<br />

largely upon <strong>the</strong> Team’s evaluation of <strong>the</strong> biology<br />

of <strong>the</strong> owl as detailed in Volume II. From <strong>the</strong>se<br />

analyses, <strong>the</strong> Team has drawn <strong>the</strong> following<br />

conclusions.<br />

<strong>Mexican</strong> spotted owls generally occupy<br />

remnants of <strong>the</strong> landscape that have experienced<br />

minimal human disturbance. We acknowledge<br />

that exceptions to this generalization occur.<br />

These remnants include inaccessible canyons,<br />

steep slopes, wilderness, and o<strong>the</strong>r environments<br />

not heavily modified by humans. Persistence of<br />

owls depends partly on <strong>the</strong>se remnant patches,<br />

but <strong>the</strong>se environments alone may be insufficient<br />

to ensure long-term conservation of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. A key point here is that<br />

not all human activies are detrimental to spotted<br />

owls. In fact, if directed appropriately, some<br />

human activities can be used to <strong>the</strong> owl’s benefit.<br />

Consequently, management must focus on<br />

75<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

creating new habitat to replace remnants that<br />

become no longer appropriate <strong>for</strong> <strong>the</strong> owl.<br />

Creation of replacement habitat hinges on<br />

understanding patterns of natural variation and<br />

modifying human activities that might conflict<br />

with <strong>the</strong> development of habitat. Natural variation<br />

across <strong>the</strong> landscape results from unique<br />

biophysical conditions at each location on <strong>the</strong><br />

land. Fur<strong>the</strong>r, effects of human activities are<br />

equally variable across <strong>the</strong> landscape. Although<br />

we cannot ascribe strict cause-effect relationships<br />

of natural processes and human activities on<br />

<strong>Mexican</strong> spotted owls, we can draw certain<br />

inferences about <strong>the</strong>ir probable impacts. The<br />

previous section detailing <strong>the</strong> conceptual framework<br />

underlying <strong>the</strong> recovery measures provides<br />

<strong>the</strong> rationale <strong>for</strong> those inferences. Thus, <strong>the</strong><br />

management recommendations (Part III) were<br />

based on two interrelated sets of in<strong>for</strong>mation: (1)<br />

basic knowledge of <strong>Mexican</strong> spotted biology;<br />

and (2) understanding how various natural<br />

processes and human activities modify <strong>the</strong><br />

environment to maintain, develop, and alter<br />

spotted owl habitat.


<strong>Recovery</strong><br />

Volume I, Part III


Removing a species or subspecies from<br />

threatened status becomes a primary management<br />

objective <strong>the</strong> moment listing is finalized.<br />

Listing a species as threatened af<strong>for</strong>ds more<br />

protection to <strong>the</strong> species than it would normally<br />

receive through o<strong>the</strong>r laws governing wildlife.<br />

Specifically, “threatened” status implies that<br />

human activities and/or natural disturbances<br />

pose greater than normal risks to <strong>the</strong> entire<br />

species or subspecies ra<strong>the</strong>r than just to individuals.<br />

Greater protection can manifest itself as<br />

more explicit and careful regulation of human<br />

activities. In some measure, a <strong>Recovery</strong> <strong>Plan</strong><br />

reconciles human needs and desires with <strong>the</strong><br />

survival needs of <strong>the</strong> threatened species or<br />

subspecies. If successful, <strong>the</strong> reconciliation<br />

process leads to an arrangement to accommodate<br />

both people and <strong>the</strong> threatened species. Ultimately,<br />

careful regulation of human activities<br />

combines with careful management of natural<br />

resources to allow removing <strong>the</strong> species from<br />

threatened status, or “delisting.” Just as listing a<br />

species requires a process of in<strong>for</strong>mation ga<strong>the</strong>ring<br />

and assessment, delisting requires a similar<br />

process.<br />

THE THE DELISTING DELISTING PROCESS<br />

PROCESS<br />

Section 4 of <strong>the</strong> Act governs <strong>the</strong> listing,<br />

delisting, and reclassification of species, <strong>the</strong><br />

designation of critical habitat, and recovery<br />

planning. Regulations implementing listing,<br />

delisting, reclassification, and critical habitat<br />

designation are codified at 50 CFR 424.<br />

The process of delisting a species or subspecies<br />

is essentially <strong>the</strong> same as that of listing: a<br />

proposed rule describing <strong>the</strong> justification <strong>for</strong> <strong>the</strong><br />

action is published in <strong>the</strong> Federal Register; a<br />

public comment period is opened, including<br />

public hearings if requested; and, within one<br />

year of <strong>the</strong> proposal, ei<strong>the</strong>r a final rule delisting<br />

<strong>the</strong> species or a notice withdrawing <strong>the</strong> proposed<br />

rule is published in <strong>the</strong> Federal Register.<br />

In considering whe<strong>the</strong>r to delist a species,<br />

<strong>the</strong> same five factors considered in <strong>the</strong> listing<br />

process (see Part I) are evaluated. While emphasis<br />

may be given to those factors leading to <strong>the</strong><br />

Volume I/Part III<br />

A. A. DELISTING<br />

DELISTING<br />

76<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

species’ listing, all of <strong>the</strong> factors must be evaluated<br />

in making a delisting determination.<br />

Section 4(c)(2) of <strong>the</strong> Act directs <strong>the</strong> FWS to<br />

conduct, at least once every five years, a review<br />

of all listed species and determine <strong>for</strong> each<br />

species whe<strong>the</strong>r it should be removed from <strong>the</strong><br />

list, reclassified from endangered to threatened,<br />

threatened to endangered, or remain in its<br />

current status. This <strong>Recovery</strong> <strong>Plan</strong> lists criteria<br />

only <strong>for</strong> delisting <strong>the</strong> <strong>Mexican</strong> spotted owl. Any<br />

decision to reclassify <strong>the</strong> subspecies to endangered<br />

status will be made by <strong>the</strong> FWS ei<strong>the</strong>r as a<br />

result of <strong>the</strong> a<strong>for</strong>ementioned mandatory review<br />

or at any o<strong>the</strong>r time in<strong>for</strong>mation becomes<br />

available indicating that reclassification is appropriate.<br />

Section 4(g) of <strong>the</strong> Act directs <strong>the</strong> FWS to<br />

implement a system in cooperation with <strong>the</strong><br />

States to monitor effectively <strong>for</strong> not less than five<br />

years <strong>the</strong> status of a species or subspecies that has<br />

been delisted due to recovery. The provisions of<br />

<strong>the</strong> Act do not apply to <strong>the</strong> delisted species<br />

during this monitoring period. However, <strong>the</strong><br />

FWS could relist a species, through <strong>the</strong> standard<br />

listing process, should monitoring indicate that<br />

<strong>the</strong> species will decline without <strong>the</strong> Act’s protection.<br />

DELISTING DELISTING CRITERIA<br />

CRITERIA<br />

We recognize that we lack data and authority<br />

to prescribe and implement monitoring strategies<br />

<strong>for</strong> Mexico. Thus, our recommendations<br />

below apply only to <strong>the</strong> U.S. range of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. We recommend that<br />

<strong>Mexican</strong> authorities develop similar delisting<br />

criteria and monitoring schemes <strong>for</strong> delisting in<br />

Mexico.<br />

Five specific criteria must be met be<strong>for</strong>e <strong>the</strong><br />

<strong>Mexican</strong> spotted owl can be delisted in <strong>the</strong> U.S.<br />

The first three criteria, which operate at a<br />

multiple-RU level, must be satisfied be<strong>for</strong>e <strong>the</strong><br />

last two criteria, which operate at <strong>the</strong> RU level,<br />

apply. These are <strong>the</strong> three overriding criteria:<br />

1. The populations in <strong>the</strong> Upper Gila<br />

Mountains, Basin and Range - East, and


Volume I/Part III<br />

Basin and Range - West RUs must be<br />

shown to be stable or increasing after 10<br />

years of monitoring, using a study design<br />

with a power of 90% to detect a 20%<br />

decline with a Type I error rate (a) of<br />

0.05.<br />

2. Scientifically-valid habitat monitoring<br />

protocols are designed and implemented<br />

to verify that (a) gross changes in<br />

macrohabitat quantity across <strong>the</strong> U.S.<br />

range of <strong>the</strong> <strong>Mexican</strong> spotted owl are<br />

stable or increasing, and (b) microhabitat<br />

modifications and trajectories within<br />

treated stands meet <strong>the</strong> intent of <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong>.<br />

3. A long-term, U.S.-rangewide management<br />

plan is in place to ensure appropriate<br />

management of <strong>the</strong> subspecies and<br />

adequate regulation of human activity<br />

over time.<br />

Once <strong>the</strong>se three criteria are satisfactorily<br />

achieved, delisting may occur in any U.S. RU<br />

that meets <strong>the</strong> final two criteria:<br />

4. Threats to <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

within <strong>the</strong> RU are sufficiently moderated<br />

and/or regulated.<br />

5. Habitat of a quality to sustain persistent<br />

<strong>Mexican</strong> spotted owl populations is<br />

stable or increasing within <strong>the</strong> RU.<br />

These criteria are, by design, redundant and<br />

dependent. Meeting one criterion, to some<br />

degree, requires meeting all or some portion of<br />

<strong>the</strong> o<strong>the</strong>r criteria. Integrating <strong>the</strong> criteria is<br />

unavoidable but never<strong>the</strong>less desirable. Progress<br />

on one translates to progress on all.<br />

Monitoring Monitoring Population Population Trends<br />

Trends<br />

For a statistically valid monitoring design, we<br />

suggest <strong>the</strong> quadrat sampling scheme described<br />

in III.C. The three RUs where population<br />

monitoring is required <strong>for</strong> delisting represent <strong>the</strong><br />

bulk of <strong>the</strong> known <strong>Mexican</strong> spotted owl population<br />

in <strong>the</strong> U.S. No population delisting criteria<br />

77<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

are applied to <strong>the</strong> remaining U.S. RUs because<br />

<strong>the</strong>y would be difficult to monitor because of<br />

<strong>the</strong> small, fragmented nature of <strong>the</strong> populations.<br />

A premise <strong>for</strong> our population monitoring<br />

approach is that <strong>the</strong> existing <strong>Mexican</strong> spotted owl<br />

population in <strong>the</strong> U.S. is adequate. This premise<br />

will be tested by monitoring population trends.<br />

If <strong>the</strong> results of monitoring indicate that <strong>the</strong><br />

U.S. population is stable or increasing over <strong>the</strong><br />

next 10 to 15 years (assuming 10 years prior to<br />

delisting followed by <strong>the</strong> required 5 years after<br />

delisting), <strong>the</strong> Team is willing to accept that <strong>the</strong><br />

current population will remain viable in <strong>the</strong><br />

<strong>for</strong>eseeable future and to assume that <strong>the</strong> population<br />

is recovered. That is, <strong>the</strong> Team believes that<br />

if <strong>the</strong> current population is able to maintain<br />

itself, or to increase, <strong>the</strong>n <strong>the</strong> population has<br />

exhibited evidence that it is of ample size to<br />

persist.<br />

Our basis <strong>for</strong> <strong>the</strong> parameters included in <strong>the</strong><br />

delisting criteria are as follows. The annual rate<br />

of change of <strong>the</strong> population within a RU can be<br />

estimated as l = N /N . A population is stable if<br />

t+1 t<br />

l = 1, decreasing if l < 1, and increasing if<br />

l > 1. A 20% reduction over a 10-year period<br />

implies a value of l = 0.978; i.e., l10 = 0.80.<br />

To conclude that a population is stable, we<br />

fail to reject <strong>the</strong> null hypo<strong>the</strong>sis that l = 1, or<br />

alternatively, that <strong>the</strong> 95% confidence interval<br />

on l includes 1. If we fail to reject this null<br />

hypo<strong>the</strong>sis, we want to ensure that <strong>the</strong> possible<br />

rate of decline is very small. Thus, we suggest a<br />

Type II error rate of 0.10, and <strong>for</strong> a 15-year<br />

period, <strong>the</strong> annual estimate of l is<br />

0.98523 = 0.8 (1/15) ^ ^ ^<br />

.<br />

For this statistical test of trend, continued<br />

persistence of <strong>the</strong> <strong>Mexican</strong> spotted owl population<br />

means <strong>the</strong> Type II error rate is more important<br />

than <strong>the</strong> Type I error rate. That is, a Type I<br />

error means that we mistakenly conclude that<br />

<strong>the</strong> population is declining when it is not.<br />

Although costly measures might be taken to<br />

reverse our incorrect perception of <strong>the</strong> trend in<br />

<strong>the</strong> owl population, <strong>the</strong> persistence of <strong>the</strong><br />

population is not threatened. In contrast, a Type<br />

II error means that we conclude <strong>the</strong> population<br />

is stable or increasing when it is really declining.<br />

Thus, persistence of <strong>the</strong> population could be in<br />

jeopardy because measures would not be taken<br />

to correct <strong>the</strong> decline. There<strong>for</strong>e, we emphasize


that a low Type II error rate of b = 0.10 (power<br />

is 1 - b = 0.90) must be met to delist <strong>the</strong> species.<br />

Several biological reasons lead us to select a<br />

time span of 10-15 years <strong>for</strong> monitoring. The<br />

mean life span (MLS) of <strong>Mexican</strong> spotted owls<br />

that reach adulthood falls within this range.<br />

MLS is calculated as 1/(-log(S)), with S representing<br />

<strong>the</strong> adult survival rate. Using S = 0.8889<br />

(SE = 0.0269), survival rates calculated from <strong>the</strong><br />

demographic study areas, <strong>the</strong> MLS is about 8.5<br />

years. Calculating confidence intervals <strong>for</strong> MLS<br />

yields 16.6 years as an upper age limit. Population<br />

turnover rates provide ano<strong>the</strong>r biological<br />

argument <strong>for</strong> <strong>the</strong> time span (x) required <strong>for</strong><br />

delisting. For example, we can estimate <strong>the</strong> time<br />

that it takes 90% of <strong>the</strong> youngest members of<br />

<strong>the</strong> adult population to completely turn over, or<br />

<strong>for</strong> 90% of <strong>the</strong> existing young adult birds to die.<br />

Given <strong>the</strong> adult S of 0.8889, solving <strong>for</strong> x in<br />

0.8889 x = 1 - 0.90 gives x = 19.6 years <strong>for</strong> 90%<br />

of a given cohort of young birds to turnover. A<br />

50% turnover would be 5.9 years, which would<br />

correspond to <strong>the</strong> median life span. For x equals<br />

10 years, 70% of <strong>the</strong> young adult population<br />

will have turned over.<br />

The time duration <strong>for</strong> <strong>the</strong> monitoring and<br />

magnitude of change required to detect a population<br />

decline are related. Thomas (1990) argued<br />

that <strong>the</strong> minimum viable population size depended<br />

on <strong>the</strong> temporal variation expected in a<br />

population. Species with much temporal variation<br />

in <strong>the</strong>ir population size might normally<br />

exhibit a 20% decline over a short period. We do<br />

not expect <strong>Mexican</strong> spotted owl populations to<br />

display much temporal variation. The most<br />

variable aspect of <strong>the</strong>ir population biology is<br />

probably recruitment, and years of little or no<br />

recruitment may occur. However, because of <strong>the</strong><br />

high adult survival rate, <strong>the</strong> decline in <strong>the</strong><br />

population during a year of no recruitment<br />

would still only be 11%. Thus, two consecutive<br />

years of no recruitment would result in a 21%<br />

decline. But <strong>the</strong> fecundity estimates presented by<br />

White et al. (1995) suggest that no recruitment<br />

is unlikely. Thus, we conclude that a 20%<br />

decline over a 10-year period indicates <strong>the</strong><br />

population is truly declining and is not <strong>the</strong> result<br />

of normal temporal variation.<br />

The choice of a Type II error rate of 0.10 is<br />

somewhat arbitrary. However, this value interacts<br />

Volume I/Part III<br />

78<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

with <strong>the</strong> choice of a 20% decline over <strong>the</strong> 10year<br />

period. Figure III.A.1 depicts a hypo<strong>the</strong>tical<br />

curve <strong>for</strong> power as a function of <strong>the</strong> size of <strong>the</strong><br />

effect being detected (labeled Detectable Effect<br />

Size in <strong>the</strong> graph). We could specify that a 15%<br />

change is detectable with a 67% power, or that a<br />

25% change is detectable with a 94% power.<br />

These statements are all equivalent in terms of<br />

<strong>the</strong> ef<strong>for</strong>t required <strong>for</strong> <strong>the</strong> monitoring protocol<br />

(as shown by <strong>the</strong> graph). This is because <strong>the</strong><br />

relationship between <strong>the</strong> detectable difference<br />

and <strong>the</strong> power to detect this difference is fixed<br />

by <strong>the</strong> monitoring ef<strong>for</strong>t (normally considered as<br />

<strong>the</strong> sample size of <strong>the</strong> statistical procedure).<br />

Thus, we have suggested that a 90% power to<br />

detect a 20% decline over 10 years is a reasonable<br />

point to fix <strong>the</strong> function that relates power<br />

and magnitude of <strong>the</strong> detectable effect.<br />

In summary, we believe 10 years is a reasonable<br />

time span <strong>for</strong> monitoring because more<br />

than half of <strong>the</strong> adult population has turned<br />

over. Fur<strong>the</strong>r, we expect that <strong>the</strong> population<br />

would have been subjected to adequate environmental<br />

variation during this 10-year period.<br />

Once <strong>the</strong> species is delisted, <strong>the</strong> additional five<br />

years of monitoring as required under <strong>the</strong> Act<br />

Figure Figure Figure III.A.1. III.A.1. Hypo<strong>the</strong>tical curve of <strong>the</strong><br />

statistical power to detect a trend in a population.


should provide fur<strong>the</strong>r assurance that <strong>the</strong> population<br />

is not declining.<br />

We believe that <strong>the</strong> delisting criterion<br />

proposed here provides positive incentives to<br />

land-management organizations to vigorously<br />

pursue <strong>the</strong> proposed population monitoring<br />

system. Delisting of <strong>the</strong> species depends on<br />

providing clearly specified evidence that <strong>the</strong><br />

population is stable or increasing. The sooner<br />

<strong>the</strong> responsible land-management organizations<br />

begin <strong>the</strong> population monitoring, <strong>the</strong> sooner <strong>the</strong><br />

owl can be delisted.<br />

O<strong>the</strong>r O<strong>the</strong>r O<strong>the</strong>r Considerations Considerations of of Population<br />

Population<br />

Monitoring<br />

Monitoring<br />

The proposed procedure <strong>for</strong> population<br />

monitoring only monitors <strong>the</strong> territorial population<br />

of owls. Because nonterritorial owls (“floaters”)<br />

do not respond to <strong>the</strong> usual methods of<br />

locating <strong>the</strong>m (i.e., calling), <strong>the</strong> only method of<br />

monitoring nonterritorial birds is via radiotracking.<br />

However, radio-tracking nonterritorial<br />

birds would require large samples of juveniles to<br />

be marked with radios, and <strong>the</strong>se radios replaced<br />

on <strong>the</strong> birds as necessary to maintain <strong>the</strong> batteries<br />

as long as <strong>the</strong> individual remained in <strong>the</strong><br />

nonterritorial population. Placing radios on<br />

spotted owls may alter <strong>the</strong>ir behavior and/or<br />

survival (Paton et al. 1991, Foster et al. 1992),<br />

making such an approach of questionable value.<br />

Thus, <strong>the</strong> Team concludes that no viable method<br />

of monitoring nonterritorial birds is available.<br />

An alternative approach to monitoring<br />

populations was considered, that of using demographic<br />

study areas. We decided against using<br />

demographic studies <strong>for</strong> three reasons.<br />

First, demographic study areas suffer from a<br />

deficiency that is not inherent in <strong>the</strong> quadrat<br />

place procedure described in III.C. Demographic<br />

study areas are chosen at <strong>the</strong> beginning<br />

of <strong>the</strong> monitoring period and must remain in<br />

place to provide appropriate data to meet <strong>the</strong>ir<br />

objectives. Because <strong>the</strong>se study areas must be<br />

permanently delimited, management practices<br />

on <strong>the</strong>m may not reflect those occurring on<br />

o<strong>the</strong>r lands. In contrast, quadrats can be randomly<br />

replaced in <strong>the</strong> sample to ensure that<br />

habitat changes and management practices<br />

Volume I/Part III<br />

79<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

adequately reflect those occurring on lands<br />

outside of <strong>the</strong> quadrats.<br />

Second, <strong>the</strong> cost of demographic study areas<br />

probably exceeds <strong>the</strong> cost of our proposed<br />

quadrat monitoring approach. The cost of<br />

conducting five demographic studies <strong>for</strong> <strong>the</strong><br />

Cali<strong>for</strong>nia spotted owl is roughly equivalent to<br />

<strong>the</strong> estimated cost of quadrat monitoring (J.<br />

Verner, FS, PSW, Fresno, pers. comm). More<br />

than five demographic study areas would be<br />

needed <strong>for</strong> a valid population monitoring<br />

scheme, thus putting <strong>the</strong> cost of demographic<br />

study areas well above <strong>the</strong> costs of <strong>the</strong> proposed<br />

quadrat sampling procedure.<br />

Finally, <strong>the</strong> two existing demographic study<br />

areas were not randomly selected from all possible<br />

demographic areas (thus not providing a<br />

defendable sample). Thus, results from <strong>the</strong><br />

existing demography studies apply only to <strong>the</strong><br />

place where <strong>the</strong>se studies were done. Fur<strong>the</strong>r,<br />

even if a demographic study approach was used,<br />

<strong>the</strong>se existing study areas may not be included in<br />

<strong>the</strong> random sample needed <strong>for</strong> a statisticallydefensible<br />

monitoring scheme.<br />

The problems outlined above with respect to<br />

using demographic studies <strong>for</strong> monitoring do<br />

not negate <strong>the</strong> usefulness of such studies. Demographic<br />

studies were designed to understand<br />

aspects of spotted owl population biology and<br />

provide a wealth of in<strong>for</strong>mation on populations,<br />

habitat characteristics, and parameters of owl<br />

fitness. They were not designed to monitor<br />

large-scale population trends.<br />

Monitoring Monitoring Monitoring Habitat Habitat Trends<br />

Trends<br />

Ganey and Dick (1995) demonstrate that<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl uses specific habitat<br />

characteristics. These features vary geographically,<br />

but within pine-oak and mixed-conifer<br />

<strong>for</strong>ests spotted owls use areas that contain large<br />

trees, snags, high log volume, multistoried stand<br />

structure, and o<strong>the</strong>r specific attributes. Presently,<br />

habitat trends <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl are<br />

unknown, and <strong>the</strong> subject of conflicting speculation<br />

(II.D). Clearly, adequate habitat of sufficient<br />

quality must exist into <strong>the</strong> future to ensure<br />

population viability. Consequently, habitat<br />

monitoring is an essential part of <strong>the</strong> recovery


process and <strong>the</strong> bird should not be delisted until<br />

monitoring can ensure unequivocally that<br />

sufficient habitat exists to support a viable<br />

population of spotted owls.<br />

Habitat monitoring should address two<br />

aspects: persistence of <strong>for</strong>est types that owls<br />

prefer (macrohabitat) and specific habitat attributes<br />

within those types (microhabitat). This<br />

roughly corresponds to <strong>the</strong> coarse and fine filters<br />

described in II.D.<br />

The first task, <strong>the</strong>n, is to quantify large-scale<br />

changes in macrohabitat across <strong>the</strong> range of <strong>the</strong><br />

bird. Given existing high fire risks and <strong>the</strong><br />

current state of southwestern <strong>for</strong>ests, we expect<br />

that net macrohabitat change over <strong>the</strong> next 10<br />

years will be negative. Although some areas will<br />

develop into habitat as <strong>the</strong> result of succession<br />

and past management activities, <strong>the</strong> Team<br />

assumes that (1) most acreage currently on a<br />

trajectory to become habitat will not do so<br />

during <strong>the</strong> life of <strong>the</strong> plan, and (2) some habitat<br />

will be lost to fire during <strong>the</strong> next 10-15 years<br />

(II.D). Thus, <strong>the</strong> Team anticipates a slight<br />

decline in <strong>the</strong> total acreage of spotted owl<br />

macrohabitat during <strong>the</strong> short term. Un<strong>for</strong>tunately,<br />

we cannot specify a priori a threshold<br />

level of habitat loss that <strong>the</strong> spotted owl population<br />

can safely endure. That relationship can<br />

only be evaluated by combining <strong>the</strong> results of<br />

population and habitat monitoring.<br />

The second task <strong>for</strong> habitat monitoring is to<br />

evaluate whe<strong>the</strong>r or not management prescriptions<br />

were implemented effectively, and whe<strong>the</strong>r<br />

treated stands will remain or become owl habitat<br />

in <strong>the</strong> near future. Prescriptions pertain primarily<br />

to <strong>the</strong> use of prescribed fire and various<br />

silvicultural tools. Monitoring to meet this<br />

objective would entail pre-treatment sampling to<br />

measure existing habitat attributes, and posttreatment<br />

sampling to verify that <strong>the</strong> prescription<br />

met <strong>the</strong> intent of <strong>the</strong> treatment. Attributes<br />

to be sampled include both those typically<br />

measured during stand examinations and also<br />

additional variables not typically measured but<br />

that are strong correlates of owl presence (e.g.,<br />

canopy cover, log volume). The general design<br />

<strong>for</strong> measuring owl habitat can be modified to<br />

monitor o<strong>the</strong>r ecosystem attributes as well. That<br />

is, additional variables can be measured besides<br />

those needed <strong>for</strong> spotted owls as required <strong>for</strong><br />

Volume I/Part III<br />

80<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

o<strong>the</strong>r ecosystem management objectives. These<br />

types of coordinated ef<strong>for</strong>ts will be crucial to<br />

meeting <strong>the</strong> monitoring needs inherent to both<br />

ecosystem and adaptive management.<br />

Long-term Long-term Management Management <strong>Plan</strong><br />

<strong>Plan</strong><br />

As described in Part I, this <strong>Recovery</strong> <strong>Plan</strong> is<br />

intended to guide management <strong>for</strong> <strong>Mexican</strong><br />

spotted owls over <strong>the</strong> next 10-15 years. If implemented<br />

as recommended, significant research<br />

will be conducted during this period and important<br />

new in<strong>for</strong>mation will become available,<br />

specifically data on owl biology, population<br />

structure, and effects of certain management<br />

practices on owl habitat. Fur<strong>the</strong>r, <strong>the</strong> guidelines<br />

that we propose <strong>for</strong> managing restricted areas<br />

(III.B) will provide a foundation upon which<br />

long-term management might be based. Evaluation<br />

of this approach and <strong>the</strong> in<strong>for</strong>mation<br />

provided through research and monitoring will<br />

be integral to developing and refining a longterm<br />

plan <strong>for</strong> managing <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl. Such a plan will be required be<strong>for</strong>e delisting<br />

can be considered.<br />

Delisting Delisting at at <strong>the</strong> <strong>the</strong> RU RU Level<br />

Level<br />

The Team recommends that once <strong>the</strong> population<br />

and habitat are shown to be stable or<br />

increasing, delisting should be considered at <strong>the</strong><br />

RU level. When delisting is considered, attention<br />

must focus on <strong>the</strong> resolution of known<br />

threats and <strong>the</strong> identification of emerging threats<br />

that could potentially compromise population<br />

viability. Similarly, spotted owl habitat must be<br />

monitored in each RU to determine trends.<br />

Monitoring will reveal habitat decline, improvement,<br />

or relative stability. The reasoning is that if<br />

<strong>the</strong> threats are removed or adequately regulated<br />

and if habitat trends are stable or showing<br />

improvement, protection under <strong>the</strong> Act will no<br />

longer be necessary. Conversely, a habitat decline<br />

or a lack of adequate regulatory mechanisms<br />

(o<strong>the</strong>r than provided by <strong>the</strong> Act) would warrant<br />

continued protection under <strong>the</strong> Act.<br />

The reasoning behind monitoring population<br />

levels within three RUs and habitat in all<br />

RUs is as follows. A viable core population will


exist in <strong>the</strong> three RUs if <strong>the</strong> population is shown<br />

to be stable or increasing. There<strong>for</strong>e, if habitat<br />

rangewide is also stable or increasing, <strong>the</strong> core<br />

population is provided <strong>the</strong> opportunity of<br />

expanding its area and greatly increasing <strong>the</strong><br />

persistence probability of <strong>the</strong> subspecies. As<br />

discussed by Keitt et al. (1995), some key<br />

unoccupied habitat patches are potentially<br />

significant in <strong>the</strong> expansion of <strong>the</strong> core population.<br />

Moderating Moderating and and Regulating Regulating Threats Threats<br />

Threats<br />

Threats to be moderated include those that<br />

need site-specific treatment to alleviate <strong>the</strong>m.<br />

The primary threat throughout <strong>the</strong> <strong>for</strong>ested U.S.<br />

range of <strong>the</strong> <strong>Mexican</strong> spotted owl is <strong>the</strong> threat of<br />

widescale, stand-replacing fire. For threats to be<br />

considered as moderated, reasonable progress<br />

must have been made in removing <strong>the</strong> threats<br />

and adequate assurance, in <strong>the</strong> <strong>for</strong>m of <strong>the</strong> longterm<br />

management plan described above, must<br />

exist that those programs will continue as necessary.<br />

Threats to be regulated include those resulting<br />

from agency management programs or o<strong>the</strong>r<br />

anthropogenic activities that are ei<strong>the</strong>r ongoing<br />

Volume I/Part III<br />

81<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

or reasonably certain to occur. A partial listing of<br />

threats besides fire include:<br />

1. Timber or fuelwood harvest that ei<strong>the</strong>r<br />

directly affects habitat within a territory<br />

or indirectly affects <strong>the</strong> owl by collateral<br />

activity adjoining owl territories;<br />

2. urban and rural land development;<br />

3. livestock and wildlife grazing;<br />

4. recreation involving both consumptive<br />

and nonconsumptive activities.<br />

Habitat Habitat Trends Trends Within Within Within <strong>Recovery</strong> <strong>Recovery</strong> Units<br />

Units<br />

For <strong>the</strong> spotted owl to be delisted within any<br />

RU, <strong>the</strong> following conditions must be met. First,<br />

threats to <strong>the</strong> continued loss of habitat and key<br />

habitat components must be moderated and<br />

regulated as detailed in <strong>the</strong> previous section.<br />

Second, habitat trends must be monitored to<br />

assess gross changes in habitat quantity within<br />

each RU. Third, effects of modifying activities<br />

within existing and potential spotted owl habitat<br />

must be monitored to ensure that existing<br />

habitat is maintained and potential habitat is<br />

progressing towards becoming replacement<br />

habitat.


The <strong>Recovery</strong> <strong>Plan</strong> recommendations are a<br />

combination of (1) protection of both occupied<br />

habitats and unoccupied areas approaching<br />

characteristics of nesting habitat, and (2) implementation<br />

of ecosystem management within<br />

unoccupied but potential habitat. The goal is to<br />

protect conditions and structures used by spotted<br />

owls where <strong>the</strong>y exist and to set o<strong>the</strong>r stands<br />

on a trajectory to grow into replacement nest<br />

habitat or to provide conditions <strong>for</strong> <strong>for</strong>aging and<br />

dispersal. By necessity this <strong>Plan</strong> is a hybrid<br />

approach because <strong>the</strong> status of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl as a threatened species requires some<br />

level of protection until <strong>the</strong> subspecies is<br />

delisted. These constraints modify ways and<br />

opportunities to manage ecosystems within<br />

landscapes where owls occur or might occur in<br />

<strong>the</strong> future. We are applying ecosystem management<br />

in two slightly different ways. Within<br />

unoccupied mixed-conifer and pine-oak <strong>for</strong>est<br />

on


number of guiding principles. These are enumerated<br />

below.<br />

Assumptions<br />

Assumptions<br />

1. <strong>Spotted</strong> owl distribution is limited<br />

primarily by <strong>the</strong> availability of habitat<br />

types used <strong>for</strong> nesting and/or roosting.<br />

2. Habitats used <strong>for</strong> nesting/roosting also<br />

provide adequate conditions <strong>for</strong> <strong>for</strong>aging<br />

and dispersal activities. Thus, providing<br />

nesting/roosting habitat partially meets<br />

o<strong>the</strong>r survival requirements as well. In<br />

turn, some stand structures not used <strong>for</strong><br />

nesting/roosting may provide adequate<br />

conditions <strong>for</strong> o<strong>the</strong>r activities such as<br />

<strong>for</strong>aging and dispersal. These include<br />

some stands in younger seral stages than<br />

typical nesting/roosting habitat.<br />

3. Nesting/roosting habitat in <strong>for</strong>est environments<br />

is typified by certain structural<br />

features, including large trees and late<br />

seral characteristics, which are common<br />

in, but not restricted to, old-growth<br />

<strong>for</strong>ests.<br />

Volume I/Part III<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure III.B.1. III.B.1. Conceptualization of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> and needs <strong>for</strong> delisting of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl depicting <strong>the</strong> interdependency of population monitoring, habitat monitoring, and management<br />

recommendations.<br />

83<br />

4. Forested nesting/roosting habitat is<br />

typically found in mixed-conifer, pineoak,<br />

and riparian <strong>for</strong>ests. O<strong>the</strong>r habitat<br />

types are used primarily <strong>for</strong> <strong>for</strong>aging,<br />

dispersal, or wintering. Thus, <strong>the</strong> distribution<br />

of nesting/roosting habitat is<br />

naturally discontinuous. Fur<strong>the</strong>r, <strong>the</strong><br />

potential distribution of such habitat is<br />

quite limited in some areas.<br />

5. The presence of shade-intolerant species<br />

in many spotted owl nest/roost stands<br />

suggests that <strong>the</strong>se areas are dynamic and<br />

have developed over time, often from<br />

more open stands. Disturbance events<br />

leading to <strong>for</strong>est canopy gaps may be<br />

important in maintaining shade-intolerant<br />

species, particularly in mixed-conifer<br />

stands.<br />

6. Existing stand structures used by <strong>Mexican</strong><br />

spotted owls <strong>for</strong> nesting/roosting<br />

generally have not been a target of<br />

planned silvicultural treatments. Where<br />

such conditions exist in managed stands,<br />

<strong>the</strong>y are more than likely an unplanned


Volume I/Part III<br />

ra<strong>the</strong>r than purposeful result. The<br />

existence of <strong>Mexican</strong> spotted owl nesting/roosting<br />

habitat usually results from<br />

<strong>the</strong> lack of recent alteration of <strong>for</strong>est<br />

structures in certain landscapes.<br />

Guiding Guiding Principles<br />

Principles<br />

1. Silvicultural applications must be evaluated<br />

over time by rigorous monitoring<br />

procedures to assess <strong>the</strong>ir effectiveness in<br />

managing or creating owl habitat.<br />

2. Obtaining large trees is a function of<br />

both time and site productivity. Similarly,<br />

many late seral characteristics<br />

typical of owl habitat, such as brokentopped<br />

trees, snags, large downed logs<br />

and <strong>the</strong> sharing of growing space among<br />

multiple shade-tolerant and intolerant<br />

species, are attained primarily through<br />

time.<br />

3. Although this <strong>Recovery</strong> <strong>Plan</strong> represents a<br />

short-term strategy, management actions<br />

recommended herein will have longterm<br />

consequences. There<strong>for</strong>e, care<br />

should be taken to preserve future<br />

options while evaluating <strong>the</strong> effectiveness<br />

of proposed treatments.<br />

GENERAL GENERAL RECOMMENDATIONS<br />

RECOMMENDATIONS<br />

General management recommendations <strong>for</strong><br />

use throughout <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl are given here. These general recommendations<br />

apply primarily to <strong>for</strong>ested areas, and<br />

aspects of <strong>the</strong> recommendations are more applicable<br />

to some locations than o<strong>the</strong>rs. Because <strong>the</strong><br />

severity of potential threats varies among RUs,<br />

<strong>the</strong> general guidelines should be prioritized and<br />

applied accordingly. Specific management<br />

priorities are emphasized in sections on individual<br />

RUs, as warranted by <strong>the</strong> differences<br />

among RUs.<br />

Three levels of habitat management are<br />

given in this <strong>Recovery</strong> <strong>Plan</strong>: protected areas,<br />

restricted areas, and o<strong>the</strong>r <strong>for</strong>est and woodland<br />

types (Figure III.B.2). Protected areas receive <strong>the</strong><br />

highest level of protection under this plan, o<strong>the</strong>r<br />

84<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>for</strong>est and woodland types <strong>the</strong> lowest. Guidelines<br />

proposed in this <strong>Recovery</strong> <strong>Plan</strong> take precedence<br />

over o<strong>the</strong>r agency management guidelines in<br />

protected areas. Guidelines <strong>for</strong> restricted areas<br />

are less specific and operate in conjunction with<br />

ecosystem management and existing management<br />

guidelines. We propose no owl-specific<br />

guidelines <strong>for</strong> lands not included in protected<br />

and restricted areas; <strong>the</strong>se areas will continue to<br />

be managed under existing guidelines, assuming<br />

that <strong>the</strong> emphasis is towards ecosystem management.<br />

One guideline that applies to all areas with<br />

any potential <strong>for</strong> owl use is to inventory <strong>for</strong><br />

spotted owls be<strong>for</strong>e implementing any management<br />

action that will alter habitat structure. If<br />

results of past inventory ef<strong>for</strong>ts can demonstrate<br />

unequivocally that no spotted owls have been<br />

detected within a given area or habitat and that<br />

<strong>the</strong> probability of detecting a bird <strong>the</strong>re is small,<br />

<strong>the</strong>n future surveys may not be needed. Under<br />

such circumstances, concurrence must be<br />

granted by <strong>the</strong> <strong>Recovery</strong> Team through <strong>the</strong><br />

appropriate RU working team.<br />

Protected Protected Areas Areas<br />

Areas<br />

Protect all <strong>Mexican</strong> spotted owl sites known<br />

from 1989 through <strong>the</strong> life of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

(Protected Activity Centers), all areas in mixedconifer<br />

and pine-oak types (defined in II.C) with<br />

slope >40% where timber harvest has not occurred<br />

in <strong>the</strong> past 20 years, and all legally and<br />

administratively reserved lands. Specific guidelines<br />

and <strong>the</strong> rationale <strong>for</strong> <strong>the</strong>se guidelines are<br />

provided below.<br />

Protected Protected Activity Activity Center Center (PAC)<br />

(PAC)<br />

Guidelines.—<br />

Guidelines.—Eight Guidelines.— specific guidelines pertain to<br />

<strong>the</strong> designation and implementation of PACs.<br />

These guidelines supersede steep slope guidelines;<br />

that is, steep slopes occurring within PACs<br />

should be managed under PAC guidelines.<br />

1. Establish PACs at all <strong>Mexican</strong> spotted<br />

owl sites known from 1989 through <strong>the</strong><br />

life of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>, including new<br />

sites located during surveys. PACs should<br />

also be established at any historical sites<br />

within <strong>the</strong> Colorado Plateau, Sou<strong>the</strong>rn<br />

Rocky Mountains - Colorado, and


Volume I/Part III<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e III.B.2. III.B.2. Generalization of protection strategies by <strong>for</strong>est/vegetation type. Proportions are not<br />

to scale.<br />

85


Volume I/Part III<br />

Sou<strong>the</strong>rn Rocky Mountains - New<br />

Mexico RUs. Identify <strong>the</strong> activity center<br />

within each PAC. “Activity center” is<br />

defined as <strong>the</strong> nest site, a roost grove<br />

commonly used during <strong>the</strong> breeding<br />

season in absence of a verified nest site,<br />

or <strong>the</strong> best roosting/nesting habitat if<br />

both nesting and roosting in<strong>for</strong>mation<br />

are lacking. Site identification should be<br />

based on <strong>the</strong> best judgement of a biologist<br />

familiar with <strong>the</strong> area. Delineate an<br />

area no less than 243 ha (600 ac) around<br />

this activity center using boundaries of<br />

known habitat polygons and/or topographic<br />

boundaries, such as ridgelines, as<br />

appropriate (Figure III.B.3). The boundary<br />

should enclose <strong>the</strong> best possible owl<br />

habitat, configured into as compact a<br />

unit as possible, with <strong>the</strong> nest or activity<br />

center located near <strong>the</strong> center. This<br />

should include as much roost/nest<br />

habitat as is reasonable, supplemented by<br />

<strong>for</strong>aging habitat where appropriate. For<br />

example, in a canyon containing mixedconifer<br />

on north-facing slopes and<br />

ponderosa pine on south-facing slopes, it<br />

may be more desirable to include some<br />

of <strong>the</strong> south-facing slopes as <strong>for</strong>aging<br />

habitat than to attempt to include 243<br />

ha (600 ac) of north-slope habitat. In<br />

many canyon situations, oval PACs may<br />

make more sense than, <strong>for</strong> example,<br />

circular PACs; but oval PACs could still<br />

include opposing canyon slopes as<br />

described above. All PACs should be<br />

retained <strong>for</strong> <strong>the</strong> life of this <strong>Recovery</strong><br />

<strong>Plan</strong>, even if spotted owls are not located<br />

<strong>the</strong>re in subsequent years. A potential<br />

exception to this rule is described in #8<br />

below. Feedback on PAC delineation<br />

should be provided to managers through<br />

RU working groups (see Part IV). PAC<br />

boundaries may not overlap.<br />

2. No harvest of trees >22.4 cm (9 in) dbh<br />

is allowed in PACs. Harvest of any trees<br />

is only permitted as it pertains to 5<br />

below.<br />

3. Fuelwood harvest within PACs should be<br />

managed in such a way as to minimize<br />

86<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

effects on <strong>the</strong> owl, its prey, and <strong>the</strong>ir<br />

habitats. The most effective management<br />

to meet <strong>the</strong>se objectives may be to<br />

prohibit such harvest. However, we<br />

recognize that it may be virtually impossible<br />

to en<strong>for</strong>ce such a prohibition and<br />

restrict access to all PACs <strong>for</strong> fuelwood<br />

harvest. When fuelwood harvest in PACs<br />

is unavoidable, we advocate <strong>the</strong> use of<br />

various <strong>for</strong>ms of management that can<br />

regulate access to PACs and to <strong>the</strong> types<br />

of fuels harvested. Potential <strong>for</strong>ms of<br />

fuelwood management include road<br />

closures, prohibiting harvest of important<br />

tree species such as oaks, prohibiting<br />

harvest of key habitat components such<br />

as snags and large downed logs (>30 cm<br />

[12 inch] midpoint diameter), and<br />

encouraging <strong>the</strong> harvest of small diameter<br />

conifers in accord with 5c below.<br />

Prohibiting fuelwood harvest of key<br />

habitat components such as oaks, snags,<br />

and large logs should be applied both<br />

inside and outside of PACs to ensure that<br />

<strong>the</strong>se special components remain on <strong>the</strong><br />

landscape.<br />

4. Road or trail building in PACs should<br />

generally be avoided but may be allowed<br />

on a case-specific basis if pressing management<br />

reasons can be demonstrated.<br />

5. Implement a program consisting of<br />

appropriate treatments to abate fire risk.<br />

The intent of this program is to assess<br />

<strong>the</strong> combined effects of thinning and fire<br />

on spotted owls and <strong>the</strong>ir habitat. The<br />

program should be structured as follows:<br />

a) Select up to 10% of <strong>the</strong> PACs within<br />

each RU that exhibit high fire risk<br />

conditions. Nest sites must be known<br />

within <strong>the</strong>se PACs. Ideally, a paired<br />

sample of PACs should be selected to<br />

serve as control areas.<br />

b) Within each selected PAC,<br />

designate 40 ha (100 acres)<br />

centered around <strong>the</strong> nest site.<br />

This nest area should include<br />

habitat that resembles <strong>the</strong> structural


Volume I/Part III<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e III.B.3. III.B.3. Examples of protected activity center (PAC) boundaries from <strong>the</strong> Lincoln National<br />

Forest. Prepared by D. Salas and L. Cole, Lincoln NF.<br />

87


Volume I/Part III<br />

and floristic characteristics of <strong>the</strong><br />

nest site. These 40 ha (100 acres) will<br />

be deferred from <strong>the</strong> treatments<br />

described below.<br />

c) Within <strong>the</strong> remaining 203 ha (500<br />

acres), combinations of thinning<br />

trees 30 cm [12<br />

inches] midpoint diameter), grasses<br />

and <strong>for</strong>bs, and shrubs. These habitat<br />

components are strong correlates of<br />

<strong>the</strong> presence of many key prey<br />

species of <strong>the</strong> owl. Emphasis of <strong>the</strong><br />

spatial configuration of treatments<br />

should be to mimic natural mosaic<br />

patterns.<br />

d) Treatments can occur only during<br />

<strong>the</strong> nonbreeding season (1 September-28<br />

February) to minimize any<br />

potential deleterious effects on <strong>the</strong><br />

owl during <strong>the</strong> breeding season.<br />

e) Following treatments to 10% of <strong>the</strong><br />

PACs, effects on <strong>the</strong> owl, prey<br />

species, and <strong>the</strong>ir habitats should be<br />

assessed. If such effects are nonnegative,<br />

an additional sample of<br />

PACs may be treated. If negative<br />

effects are detected, <strong>the</strong>se effects<br />

must be carefully evaluated. If <strong>the</strong>y<br />

can be ameliorated by modifying<br />

treatments, those modifications<br />

should occur prior to treatment of<br />

additional PACs. If not, no additional<br />

treatments should be permitted.<br />

6. Within <strong>the</strong> remaining PACs, light<br />

burning of ground fuels may be allowed<br />

within <strong>the</strong> 500 acres surrounding <strong>the</strong><br />

100-acre PAC centers (5b above), following<br />

careful review by biologists and fuels<br />

management specialists on a case-specific<br />

basis. Burns should be designed and<br />

88<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

implemented to meet <strong>the</strong> objectives<br />

noted in 5c above. Burns should be done<br />

only during <strong>the</strong> nonbreeding season<br />

(1 September-28 February).<br />

7. Within PACS treated to reduce fire risk,<br />

ei<strong>the</strong>r by <strong>the</strong> use of prescribed fire alone<br />

or in conjunction with mechanical<br />

removal of stems and ground fuels, preand<br />

post-treatment assessments (i.e.,<br />

monitoring) of habitat conditions and<br />

owl occupancy must be done. Specific<br />

habitat characteristics that should be<br />

monitored include fuel levels, canopy<br />

cover, snag basal area, volume of large<br />

logs (>30 cm [12 inch] midpoint diameter),<br />

and live tree basal area.<br />

8. If a stand-replacing fire occurs within a<br />

PAC, timber salvage plans must be<br />

evaluated on a case-specific basis. In all<br />

cases, <strong>the</strong> PAC and a buffer extending<br />

400 m from <strong>the</strong> PAC boundary must be<br />

surveyed <strong>for</strong> owls following <strong>the</strong> fire. A<br />

minimum of four visits, spaced at least<br />

one week apart, must be conducted<br />

be<strong>for</strong>e non-occupancy can be inferred. If<br />

<strong>the</strong> PAC is still occupied by owls or if<br />

owls are nearby (i.e., within 400 m of <strong>the</strong><br />

PAC boundary), <strong>the</strong>n <strong>the</strong> extent and<br />

severity of <strong>the</strong> fire should be assessed and<br />

reconfiguration of <strong>the</strong> PAC boundaries<br />

might be considered through section 7<br />

consultation. If no owls are detected,<br />

<strong>the</strong>n section 7 consultation should be<br />

used to evaluate <strong>the</strong> proposed salvage<br />

plans. If in<strong>for</strong>mal consultation cannot<br />

resolve <strong>the</strong> issue within 30 days, <strong>the</strong><br />

appropriate RU working team should be<br />

brought into <strong>the</strong> negotiations.<br />

Salvage logging within PACs should<br />

be <strong>the</strong> exception ra<strong>the</strong>r than <strong>the</strong> rule.<br />

The <strong>Recovery</strong> Team advocates <strong>the</strong><br />

general philosophy of Beschta et al.<br />

(1995) <strong>for</strong> <strong>the</strong> use of salvage logging. In<br />

particular: (1) no management activities<br />

should be undertaken that do not<br />

protect soil integrity; (2) actions should<br />

not be done that impede natural recovery<br />

of disturbed systems; and (3) salvage


Volume I/Part III<br />

activities should maintain and enhance<br />

native species and natural recovery<br />

processes. Fur<strong>the</strong>r, any salvage should<br />

leave residual snags and logs at levels and<br />

size distributions that emulate those<br />

following pre-settlement, stand-replacing<br />

fires. Scientific in<strong>for</strong>mation applicable to<br />

local conditions should be <strong>the</strong> basis <strong>for</strong><br />

determining those levels.<br />

Rationale Rationale.—The Rationale primary objective to be<br />

achieved by <strong>the</strong>se guidelines is to protect <strong>the</strong> best<br />

available habitat <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl,<br />

while maintaining sufficient flexibility <strong>for</strong> land<br />

managers to abate high fire risks and to improve<br />

habitat conditions <strong>for</strong> <strong>the</strong> owl and its prey. We<br />

assume that <strong>the</strong> best available owl habitat is that<br />

which is currently occupied by owls, or that<br />

occupied by owls in <strong>the</strong> recent past (since 1989).<br />

The median size of <strong>the</strong> adaptive kernel contour<br />

enclosing 75% of <strong>the</strong> <strong>for</strong>aging locations <strong>for</strong> 14<br />

pairs of radio-marked owls was 241 ha (595 ac).<br />

There<strong>for</strong>e, a 243 ha (600 ac) PAC should<br />

provide a reasonable amount of protected<br />

habitat and should provide <strong>for</strong> <strong>the</strong> nest site,<br />

several roost sites, and <strong>the</strong> most proximal and<br />

highly used <strong>for</strong>aging areas. We assume that<br />

existing management guidelines and those<br />

discussed below <strong>for</strong> areas outside of PACs will<br />

ensure <strong>the</strong> existence of additional habitat appropriate<br />

<strong>for</strong> <strong>for</strong>aging.<br />

The intent of <strong>the</strong>se guidelines is not to<br />

preserve <strong>the</strong>se PACs <strong>for</strong>ever, but ra<strong>the</strong>r to<br />

protect <strong>the</strong>m until it can be demonstrated that<br />

we can create replacement habitat through active<br />

management. We describe below in <strong>the</strong> section<br />

covering restricted areas <strong>the</strong> approach <strong>for</strong> managing<br />

to create replacement habitat. Once land<br />

managers demonstrate that <strong>the</strong>y can create<br />

replacement habitat, and when monitoring<br />

indicates that populations and habitats are stable<br />

or increasing, PACs could be abolished in<br />

conjunction with delisting <strong>the</strong> owl.<br />

The Team recognizes that protection status<br />

carries some risk with respect to probabilities of<br />

catastrophic fire. The reason <strong>for</strong> <strong>the</strong> proposed<br />

management within PACs is to encourage a<br />

proactive approach to reduce fuel risks and<br />

simultaneously enhance prey habitat. If <strong>the</strong>se<br />

objectives are achieved, existing owl habitat will<br />

89<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

be maintained and in some cases enhanced,<br />

while identified risks of catastrophic fire will be<br />

lessened.<br />

Salvage logging in PACs should be allowed<br />

only if sound ecological justification is provided<br />

and if <strong>the</strong> proposed actions meet <strong>the</strong> intent of<br />

this <strong>Recovery</strong> <strong>Plan</strong>, specifically to protect existing<br />

habitat and accelerate <strong>the</strong> development of<br />

replacement habitat. Fires within PACs are not<br />

necessarily bad. In many cases, patchy fires will<br />

result in habitat heterogeneity and may benefit<br />

<strong>the</strong> owl and its prey. In such cases, adjustments<br />

to PAC boundaries are probably unnecessary and<br />

salvage should not be done. Salvage should be<br />

considered in PACs only when <strong>the</strong> fire is extensive<br />

in size and results in <strong>the</strong> mortality of a<br />

substantial proportion of trees.<br />

Steep Steep Slopes Slopes Slopes (outside (outside of of PACs) PACs)<br />

PACs)<br />

Guidelines Guidelines.—Within Guidelines mixed-conifer and pineoak<br />

types, allow no harvest of trees >22.4 cm (9<br />

inches) on any slopes >40% where timber<br />

harvest has not occurred in <strong>the</strong> past 20 years.<br />

(Mixed-conifer and pine-oak types found on<br />

steep slopes that have been treated within <strong>the</strong><br />

past 20 years are managed under restricted area<br />

guidelines below). These guidelines also apply to<br />

<strong>the</strong> bottoms of steep canyons. Thinning of trees<br />

30 cm midpoint diameter), and live tree basal<br />

area.<br />

Rationale ationale ationale.—The ationale objective of prohibiting timber<br />

harvest but allowing treatment of fuels and<br />

burning is to retain additional habitat with<br />

existing conditions similar to owl nesting/<br />

roosting habitat while reducing fire risks. These


conditions appear to be found commonly in<br />

mature/old-growth stands, and such stands are<br />

now found most commonly on steep slopes<br />

because past management practices have largely<br />

occurred on slopes


succession. The landscape mosaic resulting from<br />

such an allocation should ensure adequate<br />

nesting, roosting, and <strong>for</strong>aging habitat <strong>for</strong> <strong>the</strong><br />

owl, and habitats <strong>for</strong> its variety of prey.<br />

Existing Existing Conditions<br />

Conditions<br />

Ideally, assessments of existing conditions<br />

should follow <strong>the</strong> spatial hierarchy presented by<br />

Kaufmann et al. (1994:6). At <strong>the</strong> very least,<br />

existing distributions of seral stages should be<br />

assessed at <strong>the</strong> planning level, landscape, subregional,<br />

and regional scales (sensu Kaufmann et al.<br />

1994). We recognize that in<strong>for</strong>mation may be<br />

inadequate to conduct assessments at larger<br />

spatial scales, but this constraint should be<br />

ameliorated as resource agencies continue to<br />

acquire appropriate data. Existing vegetative<br />

conditions within mature-old stands must also<br />

be assessed to determine <strong>the</strong> treatment potentials<br />

within those stands. However, given <strong>the</strong> high<br />

frequency of recent stand-altering disturbances,<br />

many areas are likely deficient in mature to oldgrowth<br />

<strong>for</strong>ests. Thus, any treatments to <strong>the</strong>se<br />

stands should be applied judiciously, if at all.<br />

Reference Reference Conditions<br />

Conditions<br />

Nesting Nesting and and roosting roosting target/threshold<br />

target/threshold<br />

conditions conditions.—Forested conditions stands used by spotted<br />

owls have certain structural features in common.<br />

These conditions do not, nor can <strong>the</strong>y, occur<br />

everywhere. For example, many south-facing<br />

slopes may never attain this type of <strong>for</strong>est structure.<br />

It is impossible <strong>for</strong> us to imagine every<br />

possible management scenario, and this limits<br />

our ability to <strong>for</strong>mulate specific guidelines that<br />

would be appropriate to all situations. Our<br />

intent here is to protect appropriate nesting<br />

habitat structure where it exists and manage<br />

o<strong>the</strong>r stands to develop <strong>the</strong> needed structure.<br />

Although our knowledge of spotted owl<br />

habitat is incomplete, nesting/roosting stands<br />

exhibit certain identifiable features, including<br />

high tree basal area, large trees, multi-storied<br />

canopy, high canopy cover, and decadence in <strong>the</strong><br />

<strong>for</strong>m of downed logs and snags (Ganey and Dick<br />

1995). Fur<strong>the</strong>r, <strong>the</strong>se stands often contain a<br />

considerable hardwood component generally<br />

provided by Gambel oak in ponderosa pine-<br />

Gambel oak <strong>for</strong>ests and by various species (e.g.,<br />

Volume I/Part III<br />

91<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

oaks, maples, box elder, aspen) in mixed-conifer<br />

<strong>for</strong>ests.<br />

We used tree basal area, large tree (>45.7 cm<br />

[18 in] dbh) density, and tree size-class distribution<br />

as <strong>the</strong> variables to define target/threshold<br />

conditions (Table III.B.1). O<strong>the</strong>r variables such<br />

as snags and downed logs are important as well.<br />

We assume that if <strong>the</strong> basal area and tree density<br />

levels given in Table III.B.1 exist, adequate<br />

amounts of snags and downed logs (and o<strong>the</strong>r<br />

habitat elements) should be present.<br />

The values provided in Table III.B.1 represent<br />

targets in that <strong>the</strong>y define <strong>the</strong> desired<br />

conditions to be achieved with time and management.<br />

They also represent threshold conditions<br />

in that <strong>the</strong>y define minimal levels that<br />

must be maintained. That is, activities can occur<br />

within stands that exceed <strong>the</strong>se conditions, but<br />

<strong>the</strong> outcome of such activities cannot lower <strong>the</strong><br />

stands below <strong>the</strong> threshold levels unless largescale<br />

ecosystem assessments demonstrate that<br />

such conditions occur in a surplus across <strong>the</strong><br />

landscape (see below). Note that all values must<br />

be met simultaneously <strong>for</strong> a stand to meet target/<br />

threshold conditions.<br />

We used two primary types of in<strong>for</strong>mation<br />

to define target/threshold conditions. First, we<br />

used quantitative descriptions of site- and standlevel<br />

habitat conditions. Second, we estimated<br />

<strong>the</strong> proportion of <strong>the</strong> landscape that could<br />

sustain those conditions through time. A similar<br />

approach was provided <strong>for</strong> managing nor<strong>the</strong>rn<br />

goshawk habitat in <strong>the</strong> southwest (Reynolds et<br />

al. 1992). Thus, our approach is not without<br />

precedence.<br />

Despite repeated attempts by <strong>the</strong> <strong>Recovery</strong><br />

Team to obtain data from land-management<br />

agencies and researchers, only limited data were<br />

available <strong>for</strong> our analyses. We used nest-site data<br />

collected by SWCA (1992) which included plot<br />

measurements centered (1) at each nest location,<br />

(2) a random location within each nest stand,<br />

and (3) a random location within a stand adjacent<br />

to <strong>the</strong> nest stand (see Ganey and Dick<br />

[1995] <strong>for</strong> more detailed in<strong>for</strong>mation). We also<br />

used FS stand inventory data provided by <strong>the</strong><br />

Coconino, Apache-Sitgreaves, and Lincoln<br />

National Forests. These data consisted of standlevel<br />

data stratified by nest, core, and territory<br />

stands. Core and territory delineations were


92<br />

Table Table III.B.1. III.B.1. Target/threshold conditions <strong>for</strong> mixed-conifer and pine-oak <strong>for</strong>ests within restricted areas. Forest types are defined in II.C.


ased on FS management guidelines <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl provided by ID No. 2 (see<br />

Part I). We had no way to assess <strong>the</strong> accuracy of<br />

<strong>the</strong> data. Different methods were used to collect<br />

<strong>the</strong> SWCA data from those used to collect <strong>the</strong><br />

FS data, thus direct comparisons are tenuous.<br />

Nei<strong>the</strong>r data set was collected specifically to<br />

address our objectives; thus, <strong>the</strong> data were less<br />

than optimal <strong>for</strong> our purposes. Fur<strong>the</strong>r, numerous<br />

people collected data and we cannot assess<br />

inter-observer variation (Block et al. 1987). As<br />

obvious as <strong>the</strong>se points may be, <strong>the</strong>y greatly<br />

limited <strong>the</strong> inferences that we could make based<br />

on our analyses. Thus, we relied on our best<br />

professional judgement to evaluate <strong>the</strong> analyses<br />

and <strong>for</strong>mulate our recommendations.<br />

We explored several analyses to derive target/<br />

threshold conditions including empirical<br />

univariate and multivariate analyses, and modeling.<br />

Most analyses converged on a set a values<br />

that were validated by existing data on spotted<br />

owl nesting habitat.<br />

We used <strong>the</strong> following approach. First, we<br />

used <strong>the</strong> SWCA (1992) data to characterize nest<br />

stands on <strong>the</strong> basis of tree basal area, density of<br />

large trees, and <strong>the</strong> distribution of stand density<br />

by size classes. Next, we used <strong>the</strong> available FS<br />

stand data to identify <strong>the</strong> percentage of <strong>the</strong><br />

contemporary landscape that simultaneously<br />

meets <strong>the</strong>se values. Third, we modeled <strong>for</strong>est<br />

stands under various post-disturbance/stand<br />

initiation conditions using <strong>the</strong> Forest Vegetation<br />

Simulator (Wykoff et al. 1982, Dixon 1991,<br />

Edminster et al. 1991). This model allowed us to<br />

predict <strong>the</strong> amount of time that a stand would<br />

be in each successional stage, including <strong>the</strong><br />

amount of time <strong>the</strong> stand would retain or exceed<br />

<strong>the</strong> characteristics required <strong>for</strong> nesting and<br />

roosting. Knowledge of how long a stand meets<br />

or exceeds target conditions was used to estimate<br />

<strong>the</strong> proportion of <strong>the</strong> landscape that should meet<br />

or exceed <strong>the</strong>se stand conditions at a given time.<br />

Analyses were conducted separately <strong>for</strong><br />

mixed-conifer and pine-oak <strong>for</strong>ests. We also<br />

provide two sets of values <strong>for</strong> mixed-conifer<br />

<strong>for</strong>est that reflect different target/threshold<br />

values which are applied to different proportions<br />

of <strong>the</strong> landscape (Table III.B.1). We reiterate<br />

that all target/threshold values must be met<br />

simultaneously. For example, within mixed-<br />

Volume I/Part III<br />

93<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

conifer <strong>for</strong>ests in all RUs except Basin and Range<br />

- East, 25% of <strong>the</strong> landscape should consist of<br />

stands that have >32 m 2 /ha (150 ft 2 /acre) of tree<br />

basal area, and include >49 trees/ha (20 trees/<br />

acre) that are >45.7 cm (18 inches) dbh. Management<br />

should strive <strong>for</strong> an even distribution of<br />

stand density across all sizes classes with no less<br />

than 10% of <strong>the</strong> distribution of stand density in<br />

each of <strong>the</strong> upper three size classes: 30.5-45.7 cm<br />

(12-18 inches), 45.7-61.0 cm (18-24 inches),<br />

and >61.0 cm (24 inches). Also, 10% of <strong>the</strong> total<br />

landscape (a subset contained within <strong>the</strong> 25%<br />

discussed above), should have >39 m 2 /ha (170<br />

ft 2 /acre) of basal area in addition to <strong>the</strong> large<br />

trees and distribution of trees by size class.<br />

Target/threshold conditions <strong>for</strong> mixed-conifer<br />

<strong>for</strong>ests in <strong>the</strong> Basin and Range - East RU differ<br />

slightly in that landscape percentages are 20%<br />

and 10%. The areal percentage <strong>for</strong> Basin and<br />

Range - East RU is lower (20% compared to<br />

25%) because of <strong>the</strong> high density of owls in <strong>the</strong><br />

Sacramento Mountains which effectively places a<br />

large proportion of <strong>the</strong> landscape in protected<br />

status. Target/threshold conditions apply to only<br />

10% of <strong>the</strong> pine-oak <strong>for</strong>est (Table III.B.1).<br />

Target/threshold conditions <strong>for</strong> pine-oak <strong>for</strong>ests<br />

also require that >4.6 m 2 /ha (20 ft 2 /acre) of oak<br />

basal area be present, and that all oaks >13 cm [5<br />

inches] dbh be retained (Table III.B.1).<br />

Coarse Coarse Filter<br />

Filter<br />

We recognize that most project planning<br />

occurs at limited spatial scales such as 4,050 ha<br />

(10,000 acre) blocks. This limited spatial scale<br />

precludes ecological assessments at larger scales.<br />

Because of this limitation, <strong>the</strong> areal percentages<br />

provided in Table III.B.1 should be regarded as<br />

minimum levels <strong>for</strong> a given planning area. If a<br />

deficit occurs within <strong>the</strong> planning area, additional<br />

stands should be identified that (1) have<br />

<strong>the</strong> site potential to reach target conditions and<br />

(2) whose current conditions most closely<br />

approach those conditions. Those stands should<br />

<strong>the</strong>n be managed to achieve target conditions as<br />

rapidly as possible. However, if <strong>the</strong> proportion of<br />

<strong>the</strong> planning area that meets target conditions is<br />

greater than <strong>the</strong> percentages in Table III.B.1,<br />

none of those stands can be lowered below<br />

threshold conditions until ecosystem assessments<br />

at larger spatial scales (landscape, subregion,


egion) demonstrate that target conditions<br />

exceed <strong>the</strong> required areal percentages (Table<br />

III.B.1) at <strong>the</strong>se larger scales. This does not<br />

preclude use of treatments to reduce fire risks or<br />

lessen insect or disease problems nor does it<br />

preclude management to meet o<strong>the</strong>r ecosystem<br />

objectives as long as stand-level conditions<br />

remain at or above <strong>the</strong> threshold values given in<br />

Table III.B.1.<br />

Fine Fine Filter Filter<br />

Filter<br />

Overriding Overriding Guidelines Guidelines.—Management Guidelines<br />

activities<br />

that influence <strong>the</strong> owl and its habitat should be<br />

conducted according to <strong>the</strong> following overriding<br />

guidelines:<br />

1. Manage mixed-conifer and pine-oak<br />

<strong>for</strong>est types to provide continuous<br />

replacement nest habitat over space<br />

and time. Treatment of a particular<br />

stand depends on its capability to attain<br />

<strong>the</strong> desired stand conditions. Target<br />

stand structure would be <strong>the</strong> described<br />

conditions <strong>for</strong> nesting and roosting<br />

habitat (Table III.B.1) but only <strong>the</strong><br />

portion of <strong>the</strong> landscape that can be<br />

sustained through time should be in<br />

that condition.<br />

2. Incorporate natural variation, such as<br />

irregular tree spacing and various stand/<br />

patch sizes, into management prescriptions<br />

and attempt to mimic natural<br />

disturbance patterns.<br />

3. Maintain all species of native vegetation<br />

in <strong>the</strong> landscape, including early seral<br />

species. To allow <strong>for</strong> variation in existing<br />

stand structures and provide species<br />

diversity, both uneven-aged and evenaged<br />

systems may be used as appropriate.<br />

4. Allow natural canopy gap processes to<br />

occur, thus producing horizontal variation<br />

in stand structure.<br />

Specific Specific Guidelines Guidelines—The Guidelines following guidelines<br />

are intended to minimize threats to <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, retain and enhance important but<br />

difficult-to-replace habitat elements, and provide<br />

management flexibility.<br />

Volume I/Part III<br />

94<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

1. Emphasis should be placed on unevenaged<br />

management systems. Existing<br />

stand conditions will determine which<br />

silvicultural system is appropriate.<br />

2. Extend rotation ages <strong>for</strong> even-aged<br />

stands to >200 years. Silvicultural prescriptions<br />

should explicitly state when<br />

vegetative manipulation will cease until<br />

rotation age is reached. This age may<br />

depend on site quality, but ceasing<br />

activity at 140 years and allowing 60<br />

years <strong>for</strong> unaltered stand maturation and<br />

senescence seems reasonable.<br />

3. Within pine-oak types, emphasis should<br />

be placed on management that retains<br />

existing large oaks and promotes <strong>the</strong><br />

growth of additional large oaks.<br />

4. Retain all trees >61 cm [24 in] dbh.<br />

5. Retain hardwoods, large down logs, large<br />

trees, and snags.<br />

6. Management priority should be placed<br />

on reducing identified risks to spotted<br />

owl habitat. The primary existing threat<br />

is catastrophic wildfire. Thus, we<br />

strongly encourage <strong>the</strong> use of prescribed<br />

and prescribed natural fire to reduce<br />

hazardous fuel accumulations. Thinning<br />

from below may be desirable or necessary<br />

be<strong>for</strong>e burning to reduce ladder fuels and<br />

<strong>the</strong> risk of crown fire. Such thinning<br />

must emphasize irregular tree spacing.<br />

7. No stand that meets threshold conditions<br />

can be treated in such a way as to<br />

lower that stand below those conditions<br />

until ecosystem assessments can document<br />

that a surplus of <strong>the</strong>se stands exist<br />

at larger landscape levels (e.g., no less<br />

than <strong>the</strong> size of a FS District). This does<br />

not preclude use of treatments to reduce<br />

fire risks or lessen insect or disease<br />

problems, nor does it preclude management<br />

to meet o<strong>the</strong>r ecosystem objectives<br />

as long as stand-level conditions remain<br />

at or above <strong>the</strong> threshold values given in<br />

Table III.B.1.


Rationale Rationale Rationale.— Rationale Rationale .— .—The .— collective goal of <strong>the</strong>se general<br />

and specific guidelines is to provide spotted owl<br />

habitat that is well distributed over space and<br />

time. To accomplish this goal requires maintaining<br />

or creating stand structures typical of nesting<br />

and roosting habitats and sustaining <strong>the</strong>m in<br />

sufficient amounts and distribution to support a<br />

healthy population of <strong>Mexican</strong> spotted owls. A<br />

few guidelines merit fur<strong>the</strong>r comment.<br />

Retaining large trees is desirable because <strong>the</strong>y<br />

are impossible to replace quickly and because<br />

<strong>the</strong>y are common features of nesting and roosting<br />

habitats <strong>for</strong> <strong>the</strong> owl. Fire, viewed as a natural<br />

<strong>for</strong>mative process ra<strong>the</strong>r than as a destructive<br />

anthropogenic process, can be used advantageously<br />

to maintain or improve spotted owl<br />

habitat.<br />

The guidelines presented above should not<br />

be misconstrued as onetime management events.<br />

For example, large trees and snags are required<br />

by <strong>the</strong> spotted owl and will continue to be<br />

needed by <strong>the</strong> owl in <strong>the</strong> future. Fur<strong>the</strong>r, <strong>the</strong><br />

approach outlined above provides a foundation<br />

<strong>for</strong> <strong>the</strong> development of a long-term management<br />

strategy. Once <strong>the</strong> owl is delisted, we expect that<br />

this general template can be evaluated, finetuned,<br />

and possibly applied to PACs and steep<br />

slopes.<br />

Riparian Riparian Communities<br />

Communities<br />

Guidelines Guidelines.—<br />

Guidelines .— .—The .— goals of <strong>the</strong>se guidelines are<br />

to maintain healthy riparian ecosystems where<br />

<strong>the</strong>y exist and initiate restoration measures to<br />

return degraded areas to healthy conditions.<br />

1. Maintain riparian broad-leaved <strong>for</strong>ests in<br />

a healthy condition where <strong>the</strong>y occur,<br />

especially in canyon-bottom situations.<br />

Where such <strong>for</strong>ests are not regenerating<br />

adequately, active management may be<br />

necessary. Possible actions to restore<br />

<strong>the</strong>se <strong>for</strong>ests may include reducing<br />

grazing pressure, establishing riparian<br />

exclosures to manage <strong>for</strong>age use better,<br />

and shifting to winter grazing seasons.<br />

2. Restore lowland riparian areas. <strong>Spotted</strong><br />

owls once nested in riparian gallery<br />

<strong>for</strong>ests. Conceivably, restored riparian<br />

<strong>for</strong>ests could contribute additional<br />

Volume I/Part III<br />

95<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

nesting habitat in <strong>the</strong> future and could<br />

also create a landscape that is more<br />

effectively connected <strong>for</strong> dispersing owls.<br />

3. Emphasize a mix of size and age classes<br />

of trees. The mix should include large<br />

mature trees, vertical diversity, and o<strong>the</strong>r<br />

structural and floristic characteristics that<br />

typify natural riparian conditions.<br />

Rationale Rationale.— Rationale .— .—We .— assume that riparian <strong>for</strong>ests<br />

provide important habitat <strong>for</strong> spotted owls.<br />

Many riparian systems within <strong>the</strong> range of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl are extremely degraded as<br />

<strong>the</strong> result of past management practices. Because<br />

many of <strong>the</strong>se systems are degraded and little<br />

documentation of recent owl use exists, we have<br />

little empirical in<strong>for</strong>mation upon which to<br />

provide specific guidelines. Thus, our underlying<br />

premise is that if riparian systems are restored to<br />

more natural conditions, <strong>the</strong> needs of <strong>the</strong> owl<br />

(and numerous o<strong>the</strong>r species) will be satisfied.<br />

This is particularly true in canyon-bottom<br />

situations at middle and lower elevations where<br />

little o<strong>the</strong>r typical nesting or roosting habitat<br />

may be available. We know that canyon bottoms<br />

are used extensively by <strong>the</strong> owl, thus it is important<br />

to preserve and increase <strong>the</strong> quality of such<br />

habitat. We anticipate that PACs will include<br />

some of <strong>the</strong> best of this type of habitat that still<br />

exists, but increasing <strong>the</strong> quantity and distribution<br />

of healthy riparian habitats provides <strong>the</strong><br />

potential <strong>for</strong> increasing spotted owl habitat.<br />

Fur<strong>the</strong>rmore, maintenance of existing healthy<br />

riparian systems and restoration of those that are<br />

degraded will benefit numerous riparian-dependent<br />

flora and fauna, and ecosystem health<br />

across <strong>the</strong> landscape.<br />

O<strong>the</strong>r O<strong>the</strong>r Forest Forest and and Woodland Woodland Types<br />

Types<br />

We propose no specific guidelines <strong>for</strong> several<br />

<strong>for</strong>est and woodland community types where<br />

<strong>the</strong>y occur outside PACs. These include ponderosa<br />

pine, spruce-fir, pinyon-juniper, and aspen<br />

as defined in II.C. We emphasize, however, that<br />

<strong>the</strong> lack of specific management guidelines<br />

within this plan does not imply that we regard<br />

<strong>the</strong>se types as unimportant to <strong>the</strong> <strong>Mexican</strong><br />

spotted owl.


The Team’s rationale <strong>for</strong> <strong>the</strong>se recommendations<br />

is based on extant in<strong>for</strong>mation on <strong>the</strong><br />

natural history of <strong>the</strong> <strong>Mexican</strong> spotted owl as<br />

summarized in Part II.A and detailed in Volume<br />

II. These <strong>for</strong>ests and woodlands are not typically<br />

used <strong>for</strong> nesting and roosting. However, <strong>the</strong>y<br />

may provide habitat <strong>for</strong> <strong>for</strong>aging and possibly <strong>for</strong><br />

both dispersing and wintering spotted owls. Our<br />

grasp of <strong>the</strong> owl’s natural history regarding <strong>the</strong>se<br />

behaviors is incomplete, so we do not fully<br />

understand <strong>the</strong> structural features <strong>the</strong> owl<br />

requires <strong>for</strong> pursuing <strong>the</strong>se activities in <strong>the</strong>se<br />

<strong>for</strong>est and woodland types. Fur<strong>the</strong>rmore, some<br />

of <strong>the</strong> best <strong>for</strong>aging habitat should be protected<br />

in PACs. All of <strong>the</strong>se circumstances allow us to<br />

be less restrictive in <strong>the</strong>se community types<br />

without harming <strong>the</strong> owl or compromising its<br />

primary habitat.<br />

With <strong>the</strong> exception of <strong>the</strong> acreage of <strong>the</strong>se<br />

types contained within PACs, we assume that<br />

<strong>the</strong> remaining lands are used primarily <strong>for</strong><br />

<strong>for</strong>aging, wintering, migration, and dispersal.<br />

Thus, we contend that existing and planned<br />

management <strong>for</strong> <strong>the</strong>se types will maintain or<br />

improve habitat <strong>for</strong> <strong>the</strong>se needs of <strong>the</strong> owl. This<br />

contention is based largely on <strong>the</strong> assumption<br />

that existing old-growth areas will be maintained<br />

across <strong>the</strong> landscape, silvicultural practices will<br />

favor selection over regeneration cuts, and<br />

management will be guided by ecosystem<br />

approaches that strive to provide sustainable<br />

conditions across <strong>the</strong> landscape that fall within<br />

<strong>the</strong> natural range of variation.<br />

Guidelines developed <strong>for</strong> protected and<br />

restricted areas may have useful applications<br />

when judiciously administered in <strong>the</strong>se o<strong>the</strong>r<br />

<strong>for</strong>est and woodland types. Such guidelines<br />

include managing <strong>for</strong> landscape diversity, mimicking<br />

natural disturbance patterns, incorporating<br />

natural variation in stand conditions, retaining<br />

special features such as snags and large trees,<br />

and utilizing fires as appropriate. We also emphasize<br />

<strong>the</strong> need <strong>for</strong> proactive fuels management<br />

where appropriate. Decreasing fire risks within<br />

<strong>the</strong>se types, particularly ponderosa pine <strong>for</strong>ests,<br />

will also decrease fire risks to adjoining protected<br />

and restricted areas by minimizing <strong>the</strong> probability<br />

of large landscape-level crown fires that could<br />

impinge upon occupied or potential nesting<br />

habitat.<br />

Volume I/Part III<br />

96<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

GRAZING GRAZING RECOMMENDATIONS<br />

RECOMMENDATIONS<br />

The explicit goals of managing grazing in<br />

spotted owl habitat reflect <strong>the</strong> four manifested<br />

influences of grazing discussed in II.D. Those<br />

influences were (1) altered prey availability, (2)<br />

altered susceptibility to fire, (3) degeneration of<br />

riparian plant communities, and (4) impaired<br />

ability of plant communities to develop into<br />

spotted owl habitat. The goals <strong>the</strong>n become (1)<br />

to maintain or enhance prey availability, (2) to<br />

maintain potential <strong>for</strong> beneficial ground fires<br />

while inhibiting potential <strong>for</strong> destructive standreplacing<br />

fire, (3) to promote natural and<br />

healthy riparian plant communities, and (4) to<br />

preserve <strong>the</strong> processes that ultimately develop<br />

spotted owl habitat.<br />

The Team strongly advocates field monitoring<br />

and experimental research related to <strong>the</strong><br />

impacts of grazing on <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Only through monitoring and research can we<br />

(1) develop a comprehensive understanding of<br />

how grazing affects <strong>the</strong> habitat of <strong>the</strong> owl and its<br />

prey; (2) determine <strong>the</strong> effectiveness of current<br />

grazing standards and guidelines as <strong>the</strong>y relate to<br />

<strong>the</strong> owl’s needs; and (3) devise grazing strategies<br />

that can benefit <strong>the</strong> owl and its prey.<br />

Grazing Grazing Guidelines Guidelines<br />

Guidelines<br />

The following guidelines should be applied<br />

to all protected and restricted areas:<br />

1. Monitor grazing use by livestock and<br />

wildlife in “key grazing areas.” Key<br />

grazing areas are primarily riparian areas,<br />

meadows, and oak types. Monitoring<br />

should begin by determining current<br />

levels of use plus current composition,<br />

density, and vigor of <strong>the</strong> plants. Ultimately,<br />

monitoring should detect any<br />

change in <strong>the</strong> relative composition of<br />

herbaceous and woody plants. The intent<br />

is to maintain good to excellent range<br />

conditions in key areas while accommodating<br />

<strong>the</strong> needs of <strong>the</strong> owl and its prey.


2. Implement and en<strong>for</strong>ce grazing utilization<br />

standards that would attain good to<br />

excellent range conditions within <strong>the</strong> key<br />

grazing areas. Use standards (e.g., FS<br />

Region 3, Range Analysis Handbook)<br />

that have been developed <strong>for</strong> local<br />

geographic areas and habitat types-particularly<br />

in key habitats such as<br />

riparian areas, meadows, and oak types-that<br />

incorporate allowable use levels<br />

based on current range condition, key<br />

species, and <strong>the</strong> type of grazing system.<br />

Establish maximum allowable use levels<br />

that are conservative and that will<br />

expedite attaining and maintaining good<br />

to excellent range conditions. The<br />

purpose of establishing <strong>the</strong>se use levels is<br />

to ensure allowable use of plant species<br />

to maintain plant diversity, density,<br />

vigor, and regeneration over time.<br />

Additionally, a primary purpose is to<br />

maintain or restore adequate levels of<br />

residual plant cover, fruits, seeds, and<br />

regeneration to provide <strong>for</strong> <strong>the</strong> needs of<br />

prey species and development of future<br />

owl <strong>for</strong>aging and dispersal habitat.<br />

3. Implement management strategies that<br />

will restore good conditions to degraded<br />

riparian communities as soon as possible.<br />

Strategies may include reductions in<br />

grazing levels and increased numbers of<br />

exclosures (i.e., fencing) to protect<br />

riparian plant cover and regeneration,<br />

and to prevent damage to stream banks<br />

and channels (Clary and Webster 1989,<br />

Platts 1990). In many cases, degraded<br />

riparian areas may require complete rest<br />

<strong>for</strong> periods from a few years to 15 years<br />

<strong>for</strong> <strong>the</strong> area to recover (Kennedy 1977,<br />

Rickard and Cushing 1982, Clary and<br />

Webster 1989). Additional strategies may<br />

include <strong>the</strong> use of riparian pastures,<br />

limited winter use, double rest-rotation,<br />

and o<strong>the</strong>r methods that emphasize<br />

riparian vegetation and stream bank/<br />

channel recovery (Platts 1990). Riparian<br />

restoration projects that include <strong>the</strong> use<br />

of exclosures need not require exclosures<br />

along <strong>the</strong> entire drainage course at one<br />

Volume I/Part III<br />

97<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

time. Ra<strong>the</strong>r, systematic use of exclosures<br />

that protect <strong>the</strong> most sensitive portions<br />

of riparian habitats is encouraged.<br />

Riparian areas can also benefit from<br />

protection of adjacent upland areas<br />

(Bryant 1982). Placement of exclosures<br />

(controls) and areas open to grazing<br />

(treatments) should be designed to<br />

permit determination of effects on<br />

several ecological responses (e.g., vegetation,<br />

erosion, water quality, prey availability).<br />

Rationale Rationale <strong>for</strong> <strong>for</strong> <strong>for</strong> Grazing Grazing Guidelines<br />

Guidelines<br />

Some effects of excessive grazing on vegetation<br />

and habitat features are predictably negative,<br />

particularly in riparian communities.<br />

However, <strong>the</strong> collective effects of grazing are<br />

nei<strong>the</strong>r always predictable nor always negative.<br />

Effects depend on site-specific factors such as <strong>the</strong><br />

grazing system, condition of <strong>the</strong> plant community<br />

prior to livestock grazing, soil types, climate,<br />

community composition of plant species, and<br />

<strong>the</strong> presence or absence of aggressive exotic plant<br />

species. Succinctly, predictability is inexact; and<br />

without predictability <strong>the</strong> Team cannot give<br />

detailed and specific recommendations.<br />

We suggest that, when implemented and<br />

en<strong>for</strong>ced, general guidelines and <strong>the</strong> standards<br />

<strong>the</strong>y prescribe will promote and maintain good<br />

to excellent range conditions over time and<br />

across communities used by <strong>the</strong> owl. Despite our<br />

imprecise knowledge of how grazing affects<br />

spotted owl habitat, <strong>the</strong> collective body of<br />

general knowledge regarding <strong>the</strong> impacts of<br />

grazing on wildlife mandates prudence. The<br />

Team believes that understanding how grazing<br />

affects <strong>the</strong> owl is paramount, and we strongly<br />

urge that specific grazing practices and levels of<br />

grazing use be carefully evaluated through an<br />

experimental approach (Bock et al. 1993).<br />

Habitats in protected and restricted areas<br />

should receive high-priority management attention<br />

relative to grazing. Any riparian communities<br />

of potential importance <strong>for</strong> spotted owl<br />

dispersal and wintering habitat should also<br />

receive high-priority attention. Such attention<br />

will not only benefit spotted owls but many<br />

o<strong>the</strong>r species in <strong>the</strong> Southwest as well (Hubbard


1977). Fundamental to <strong>the</strong> guidelines <strong>for</strong><br />

grazing is <strong>the</strong> assumption that individual actions<br />

have collective effects. For example, <strong>the</strong> shortterm<br />

goal of exclusion by fencing is to protect<br />

riparian plants and to prevent physical damage<br />

to stream banks and channels (Clary and<br />

Webster 1989, Platts 1990). The long-term goal<br />

is that such short-term protection ultimately<br />

allows spotted owl habitat to develop. Implicit in<br />

this rationale is that excessive grazing sustained<br />

<strong>for</strong> long periods not only deteriorates potential<br />

or actual spotted owl habitat but it also inevitably<br />

leads to a deterioration of <strong>the</strong> very qualities<br />

that make an area attractive <strong>for</strong> grazing in <strong>the</strong><br />

first place.<br />

We also assert that attainment and maintenance<br />

of good to excellent conditions in key<br />

grazing areas will translate to better conditions in<br />

<strong>the</strong> uplands. Most native and exotic ungulates<br />

preferentially graze within key areas such as<br />

meadows and riparian areas. We assume that if<br />

<strong>the</strong>se key areas exhibit ecologically good conditions,<br />

upland <strong>for</strong>ests and woodlands should also<br />

be in good condition. Thus, negative effects of<br />

grazing that lead to <strong>the</strong> establishment of ladder<br />

fuels and “dog-hair” thickets may be ameliorated.<br />

RECREATION<br />

RECREATION<br />

RECOMMENDATIONS<br />

RECOMMENDATIONS<br />

Guidelines<br />

Guidelines<br />

The following guidelines should be applied<br />

to all protected and restricted areas:<br />

1. No construction, ei<strong>the</strong>r of new facilities<br />

or <strong>for</strong> expanding existing facilities,<br />

should take place within PACs during<br />

<strong>the</strong> breeding season, 1 March through<br />

31 August. Any construction within<br />

PACs during <strong>the</strong> nonbreeding season<br />

should be considered on a case-specific<br />

basis. Modifications to existing facilities<br />

pertaining to public safety and routine<br />

maintenance are excepted.<br />

2. Managers should, on a case-specific basis,<br />

assess <strong>the</strong> presence and intensity of<br />

allowable recreational activities within<br />

Volume I/Part III<br />

98<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

PACs. Spatial and temporal restrictions<br />

should be considered <strong>for</strong> new activities.<br />

3. Seasonal closures of specifically designated<br />

recreational activities should be<br />

considered where appropriate.<br />

RECOVERY RECOVERY UNIT<br />

UNIT<br />

CONSIDERATIONS<br />

CONSIDERATIONS<br />

We review below primary threats within each<br />

recovery unit. Some threats are ubiquitous across<br />

<strong>the</strong> range of <strong>the</strong> spotted owl, whereas o<strong>the</strong>rs are<br />

limited to one or few RUs. To place <strong>the</strong>se threats<br />

in perspective, we review below relevant in<strong>for</strong>mation<br />

from Part II.B <strong>for</strong> each RU. Management<br />

priorities within each RU should focus on<br />

<strong>the</strong> threats identified below.<br />

One management consideration that applies<br />

to all RUs is <strong>the</strong> potential <strong>for</strong> migration and<br />

dispersal of spotted owls within and among RUs.<br />

Admittedly, we know very little of <strong>the</strong> prevalence<br />

of such movements, nor do we know much of<br />

<strong>the</strong> habitats used. We suspect, however, that<br />

movements of birds may be important to gene<br />

flow and <strong>the</strong> maintenance of a metapopulation<br />

structure (Keitt et al. 1995). Thus, ef<strong>for</strong>ts should<br />

be made to preserve options by maintaining and<br />

enhancing potential avenues <strong>for</strong> migration and<br />

dispersal. This could be particularly important in<br />

specific canyons, riparian areas, and mountain<br />

ranges that might provide links within, between,<br />

and among RUs.<br />

Colorado Colorado Colorado Plateau Plateau<br />

Plateau<br />

The Colorado Plateau is <strong>the</strong> largest of all<br />

U.S. RUs. It encompasses <strong>the</strong> sou<strong>the</strong>rn half of<br />

Utah, much of nor<strong>the</strong>rn Arizona, most of<br />

northwestern New Mexico, and a small portion<br />

of southwestern Colorado. In <strong>the</strong> northwestern<br />

portion of this RU, owls have been located in<br />

steep-walled canyons with apparent concentrations<br />

in <strong>the</strong> areas of Zion N.P., Capitol Reef<br />

N.P., western Abajo Mountains, and<br />

Canyonlands N.P. Historical records are available<br />

from <strong>for</strong>ested habitats on <strong>the</strong> Kaibab Plateau of<br />

nor<strong>the</strong>rn Arizona. In <strong>the</strong> sou<strong>the</strong>astern portion of<br />

this RU, owls occur in both steep-sloped, mixedconifer<br />

<strong>for</strong>ested canyons and steep-walled


canyons on <strong>the</strong> Navajo Indian Reservation;<br />

known concentrations occur in <strong>the</strong> Black Mesa<br />

area and <strong>the</strong> Chuska Mountains. <strong>Owl</strong>s have also<br />

been located in mixed-conifer habitats in <strong>the</strong><br />

Zuni Mountains and on Mount Taylor (Figure<br />

II.B.3).<br />

<strong>Owl</strong> distribution in this RU appears to be<br />

highly fragmented. This distributional pattern<br />

may be natural or <strong>the</strong> result of inadequate survey<br />

ef<strong>for</strong>t in some parts of <strong>the</strong> RU. Extensive surveys<br />

have, however, been completed in <strong>the</strong> sou<strong>the</strong>rn<br />

Utah portion of this RU. Here, breeding owls<br />

have been found only in canyons where <strong>the</strong>y<br />

nest and roost in caves and on ledges. In sou<strong>the</strong>rn<br />

Utah, no breeding owls have been located in<br />

hundreds of thousands of hectares surveyed in<br />

mixed-conifer or o<strong>the</strong>r <strong>for</strong>est types in areas with<br />

less than 40% slope. There<strong>for</strong>e, we recommend<br />

that surveys in sou<strong>the</strong>rn Utah emphasize steep<br />

slopes and rocky canyons.<br />

Potential Potential Threats<br />

Threats<br />

Levels of recreational activity are high and<br />

increasing in some areas of this RU, such as<br />

sou<strong>the</strong>rn Utah. Some activities may potentially<br />

lead to habitat alteration or direct disturbance of<br />

owls. Fur<strong>the</strong>rmore, owls in sou<strong>the</strong>rn Utah nest<br />

and roost in canyons, and <strong>the</strong> physical structure<br />

of canyons tend to magnify disturbances and<br />

limit escape/avoidance routes <strong>for</strong> owls. Potential<br />

threats listed in order of severity <strong>for</strong> <strong>the</strong> northwestern<br />

portion of this RU include recreation,<br />

overgrazing, and road development within<br />

canyons, and catastrophic fire and timber harvest<br />

within upland <strong>for</strong>ests potentially used <strong>for</strong> <strong>for</strong>aging,<br />

dispersal, and wintering. For <strong>the</strong> sou<strong>the</strong>astern<br />

portion, threats include timber harvest,<br />

overgrazing, catastrophic fire, oil, gas, and<br />

mining development, and recreation (Utah<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> Technical Team 1994).<br />

Sou<strong>the</strong>rn Sou<strong>the</strong>rn Rocky Rocky Mountains Mountains - - Colorado<br />

Colorado<br />

Lying completely within <strong>the</strong> State of Colorado,<br />

<strong>the</strong> Sou<strong>the</strong>rn Rocky Mountains - Colorado<br />

represents <strong>the</strong> nor<strong>the</strong>astern extreme of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl’s range. Few owls have been<br />

detected in this portion of its range, and its<br />

natural history in Colorado is poorly understood.<br />

However, <strong>the</strong> condition of a species being<br />

Volume I/Part III<br />

99<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

scarce at <strong>the</strong> periphery of its range is not unusual.<br />

Nesting and roosting habitat may be<br />

primary concerns, but wintering habitat may be<br />

an important factor in this RU. Although very<br />

little is known about wintering habitat, some<br />

data suggest that birds may winter at lower<br />

elevations that include a wider range of conditions<br />

than found in breeding habitats.<br />

Potential Potential Threats<br />

Threats<br />

In order of severity, potential threats <strong>for</strong> <strong>the</strong><br />

Sou<strong>the</strong>rn Rocky Mountains - Colorado RU are<br />

catastrophic fire, recreation, urbanization, timber<br />

harvest, and road construction. Less severe<br />

threats include land exchange, oil and gas<br />

leasing, mineral development, and grazing.<br />

Singly, <strong>the</strong>se factors may have low impact, but<br />

high synergistic consequences. For example,<br />

much of <strong>the</strong> urban development in Sou<strong>the</strong>rn<br />

Rocky Mountains - Colorado currently occurs at<br />

elevations lower than those occupied by breeding<br />

owls; but development increases recreational<br />

access to public lands. Road construction or<br />

expansion causes initial disturbance, recreation<br />

facilities extend <strong>the</strong> disturbance, and <strong>the</strong><br />

improved access increases <strong>the</strong> contact between<br />

people and spotted owls. The initial activity<br />

may directly affect wintering habitat. The<br />

development threat is considered to be of low<br />

to moderate severity and is highest along <strong>the</strong><br />

Front Range.<br />

Sou<strong>the</strong>rn Sou<strong>the</strong>rn Rocky Rocky Mountains Mountains -<br />

-<br />

New New Mexico<br />

Mexico<br />

Ranking as <strong>the</strong> smallest U.S. RU, <strong>the</strong> Sou<strong>the</strong>rn<br />

Rocky Mountains - New Mexico supports<br />

one of <strong>the</strong> smallest known populations of<br />

<strong>Mexican</strong> spotted owls as well. Existing data are<br />

too incomplete to cite even a crude estimate of<br />

<strong>the</strong> RU’s spotted owl population or its density.<br />

The inability to provide crude population<br />

estimates may be partially related to <strong>the</strong> inadequacy<br />

of existing survey protocols. <strong>Owl</strong> survey<br />

crews, following <strong>the</strong> FS Region 3 survey protocol,<br />

have speculated that spotted owls in <strong>the</strong> area<br />

may not respond to calling surveys as predictably<br />

as <strong>the</strong>y do in o<strong>the</strong>r RUs. They base this opinion<br />

on <strong>the</strong> lack of response to nighttime calling in<br />

areas where owls or young have been observed


during subsequent daytime visits. Given that<br />

surveys are an important step in designating<br />

PACs and <strong>the</strong> types of management permissible,<br />

it is critical that surveys have a high probability<br />

of detecting owls. Fur<strong>the</strong>r, an ineffective, or<br />

partially effective, survey protocol leaves <strong>the</strong> FS<br />

and o<strong>the</strong>r agencies ill-equipped to manage<br />

potential threats to <strong>the</strong> <strong>Mexican</strong> spotted owl in<br />

<strong>the</strong> RU.<br />

Recent survey ef<strong>for</strong>ts on <strong>the</strong> Santa Fe National<br />

Forest suggest that some areas <strong>for</strong>merly<br />

occupied by owls appear vacant now despite <strong>the</strong><br />

fact that <strong>the</strong> habitat has not been altered appreciably<br />

(T. Johnson, Las Alamos, NM, pers.<br />

comm.). This perceived, but unconfirmed,<br />

population decline indicates <strong>the</strong> importance of<br />

protecting unoccupied habitat.<br />

Potential Potential Threats<br />

Threats<br />

The most serious threat to spotted owls in<br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico RU is<br />

wildfire and, in localized areas, timber harvest.<br />

Fire may not be as serious in canyon systems as it<br />

is in o<strong>the</strong>r areas because <strong>the</strong> open structure of<br />

steep-slope woodlands associated with canyons<br />

are not conducive to conflagration. However,<br />

dense mixed-conifer and ponderosa pine <strong>for</strong>ests<br />

outside of canyons may present <strong>the</strong> greatest fire<br />

hazards. Although <strong>the</strong>se areas may not contain<br />

owls, fires initiated in <strong>the</strong>se <strong>for</strong>ests may continue<br />

into <strong>the</strong> <strong>for</strong>ested canyon habitats. Personnel at<br />

Santa Fe National Forest have instituted an<br />

aggressive prescribed fire program in <strong>the</strong> Jemez<br />

Mountains as a way to reduce <strong>the</strong> risk of extensive<br />

fire. Though useful, prescribed fire should<br />

be used conservatively in spotted owl habitat.<br />

Timber harvest levels appear greatest on <strong>the</strong><br />

Carson National Forest where few spotted owls<br />

have been confirmed. Timber harvest levels on<br />

Santa Fe National Forest have been reduced in<br />

areas where spotted owls are known to occur.<br />

Isolation of spotted owl pairs and small populations<br />

distributed over large areas of fragmented<br />

landscape prompt concern because if <strong>the</strong>y are<br />

lost, <strong>the</strong> species disappears from entire landscapes<br />

it once inhabited. Since <strong>the</strong> spotted owl<br />

was listed, planned timber sales have mostly<br />

avoided <strong>the</strong> owl’s habitat. However, in <strong>the</strong><br />

Vallecitos Federal Sustained Yield Unit in Carson<br />

Volume I/Part III<br />

100<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

National Forest, sales are still being planned in<br />

potential spotted owl habitat despite unconfirmed<br />

sightings of spotted owls in 1993.<br />

Lesser threats to <strong>the</strong> spotted owl include<br />

human activities that produce extremely localized<br />

effects, but that may ultimately prove to<br />

have a large collective impact. Among <strong>the</strong>m are<br />

unregulated fuelwood harvesting, grazing (particularly<br />

in riparian areas), and recreation developments<br />

at ski areas. All of <strong>the</strong>se activities have<br />

<strong>the</strong> potential to degrade spotted owl habitat<br />

including habitat <strong>for</strong> <strong>the</strong> owl’s prey.<br />

Upper Upper Gila Gila Mountains<br />

Mountains<br />

The Upper Gila Mountains RU contains <strong>the</strong><br />

largest known number of <strong>Mexican</strong> spotted owls<br />

with approximately 55% of known spotted owl<br />

territories (Ward et al. 1995). The owl also<br />

appears to be more continuously distributed<br />

across this RU than any o<strong>the</strong>r (II.B). The apparent<br />

gap in owl distribution in <strong>the</strong> center of<br />

Figure II.B.6 reflects incomplete in<strong>for</strong>mation<br />

from Tribal lands ra<strong>the</strong>r than an actual discontinuity<br />

within <strong>the</strong> owl’s range.<br />

Potential Potential Threats<br />

Threats<br />

<strong>Spotted</strong> owls throughout <strong>the</strong> RU are found<br />

primarily in mixed-conifer and pine-oak <strong>for</strong>ests<br />

(Ganey and Dick 1995), often in conjunction<br />

with canyon terrain. The primary threats to<br />

spotted owls and <strong>the</strong>ir habitat are timber harvest<br />

and catastrophic fire, not necessarily in that<br />

order. Both threats could destroy <strong>for</strong>est habitat<br />

with <strong>the</strong> structural features used by spotted owls,<br />

and both could operate over large spatial scales.<br />

O<strong>the</strong>r threats within this RU include indiscriminate<br />

fuelwood cutting and overgrazing by<br />

both wildlife and livestock. These threats are not<br />

as widespread or severe as <strong>the</strong> threats discussed<br />

above, but <strong>the</strong>y can be significant in some areas.<br />

Fuelwood cutting is a problem in some areas<br />

primarily because people remove (usually illegally)<br />

large oaks. These trees appear to be critical<br />

to owls in some areas or habitats, particularly <strong>the</strong><br />

pine-oak type (Ganey et al. 1992). Fuelwood<br />

harvest can also result in loss of large snags and<br />

down logs. Both of <strong>the</strong>se habitat components are<br />

also apparently important to <strong>the</strong> owl, ei<strong>the</strong>r<br />

directly or indirectly through effects on <strong>the</strong> prey


ase (Ganey and Dick 1995, Ward and Block<br />

1995).<br />

Overgrazing is suspected to be detrimental in<br />

some areas and can affect both habitat structure<br />

and <strong>the</strong> prey base. Effects on <strong>the</strong> prey base are<br />

difficult to quantify, but removal of herbaceous<br />

vegetation can reduce both food and cover<br />

available to small mammals (Ward and Block<br />

1995). This may be especially true with respect<br />

to voles, which are often associated with dense<br />

grass cover. Direct effects on habitat are obvious<br />

in some places, particularly with respect to<br />

browsing on young Gambel oak. In some areas,<br />

oak is regenerating well but unable to grow<br />

beyond <strong>the</strong> sapling stage because of this browsing.<br />

Coupled with loss of large oaks to fuelwood<br />

harvest, maintenance of most oak stems in a<br />

sapling stage suggests a very real possibility that<br />

large oak trees will not be replaced over large<br />

areas, resulting in <strong>the</strong> loss of an important<br />

habitat component. Grazing effects on habitat<br />

are also potentially significant in canyon-bottom<br />

riparian areas. We do not attribute <strong>the</strong>se effects<br />

solely to livestock. Forage resources are shared by<br />

livestock and wild ungulates, and reducing<br />

numbers of both will likely be necessary to bring<br />

<strong>for</strong>age use to reasonable levels.<br />

Basin Basin Basin and and Range Range - - West<br />

West<br />

Sprawling across sou<strong>the</strong>rn Arizona and<br />

extreme southwestern New Mexico, <strong>the</strong> Basin<br />

and Range - West RU ranks as <strong>the</strong> second largest<br />

RU in <strong>the</strong> United States. Though it probably<br />

does not support as large a spotted owl population<br />

as <strong>the</strong> Upper Gila Mountains RU, <strong>the</strong><br />

known population ranks third highest in <strong>the</strong><br />

United States despite limited survey ef<strong>for</strong>ts in<br />

many areas. There<strong>for</strong>e, <strong>the</strong> Team regards <strong>the</strong><br />

Basin and Range - West RU as an important unit<br />

<strong>for</strong> <strong>the</strong> recovery ef<strong>for</strong>t.<br />

Potential Potential Threats<br />

Threats<br />

The Team perceives limited threats overall to<br />

spotted owls in <strong>the</strong> Basin and Range - West RU<br />

as <strong>the</strong> result of human activities. Very little<br />

timber harvest occurs in this RU, though some<br />

timber is cut in <strong>the</strong> Bradshaw Mountains of <strong>the</strong><br />

Prescott National Forest and on <strong>the</strong> San Carlos<br />

Volume I/Part III<br />

101<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Apache Reservation. The primary threats to<br />

spotted owls within this RU are catastrophic<br />

wildfire, recreation, and grazing. We detail below<br />

<strong>the</strong> nature and extent of <strong>the</strong>se and o<strong>the</strong>r potential<br />

threats.<br />

Historical ef<strong>for</strong>ts to suppress fire have<br />

allowed fuel loads to accumulate to dangerous<br />

levels within most of <strong>the</strong> wooded and <strong>for</strong>ested<br />

vegetation types in this RU. For example, <strong>the</strong><br />

1983 fire in <strong>the</strong> Animas Mountains removed <strong>the</strong><br />

coniferous <strong>for</strong>est from <strong>the</strong> higher elevations.<br />

Recent wildfires in <strong>the</strong> Pinaleno, Rincon,<br />

Chiricahua, and Huachuca Mountains also attest<br />

to <strong>the</strong> volatile situation in this region. We view<br />

<strong>the</strong> potential <strong>for</strong> catastrophic wildfire as <strong>the</strong><br />

primary threat to spotted owls in <strong>the</strong> Basin and<br />

Range - West RU.<br />

Many mountain ranges in <strong>the</strong> Coronado,<br />

Prescott, and Tonto National Forests are used<br />

heavily <strong>for</strong> recreation. This is partly because of<br />

<strong>the</strong>ir proximity to large urban areas (Tucson and<br />

Phoenix) and partly because of <strong>the</strong>ir international<br />

reputation <strong>for</strong> exceptional birding. Effects<br />

of recreation include development of roads,<br />

campgrounds, and trails, and also extraordinary<br />

use of those facilities. For example, a number of<br />

areas within <strong>the</strong> Coronado National Forest (e.g.,<br />

Madera Canyon in <strong>the</strong> Santa Ritas, Garden and<br />

Ramsey Canyons in <strong>the</strong> Huachucas, and <strong>the</strong><br />

South Fork of Cave Creek in <strong>the</strong> Chiricahuas)<br />

are world renowned <strong>for</strong> birding and receive<br />

thousands of visitors per year. Scheelite Canyon<br />

on <strong>the</strong> Fort Huachuca Army Base is visited often<br />

by birders specifically to view <strong>the</strong> pair of spotted<br />

owls that occur <strong>the</strong>re. The <strong>Mexican</strong> spotted owl,<br />

in fact, is one of <strong>the</strong> more popular species sought<br />

by birders in this region.<br />

Cattle grazing occurs throughout <strong>the</strong> RU.<br />

Impacts are greatest in <strong>the</strong> high desert grasslands,<br />

desert scrub, and riparian habitats found between<br />

mountain ranges. Moderate grazing<br />

pressures occur within mid-elevational encinal<br />

and pinyon-juniper woodlands; and grazing<br />

pressures are evident within higher elevational<br />

canyon stringers of pine-oak, mixed-conifer, and<br />

riparian <strong>for</strong>ests. Perhaps <strong>the</strong> primary threat of<br />

grazing is to <strong>the</strong> low-elevation riparian <strong>for</strong>ests.<br />

These <strong>for</strong>ests may represent critical linkages<br />

among <strong>the</strong> mountain ranges. Modified and<br />

degraded riparian <strong>for</strong>ests may inhibit dispersal


among mountain ranges and gene flow among<br />

owl subpopulations.<br />

Land ownership within <strong>the</strong> Basin and Range<br />

- West is a mosaic of public and private lands<br />

(II.B). Most major mountain ranges fall under<br />

Federal jurisdiction, with some private<br />

inholdings and o<strong>the</strong>r lands administered by <strong>the</strong><br />

San Carlos Apache and White Mountain Apache<br />

tribes. Much of <strong>the</strong> shrublands and grasslands<br />

between mountain ranges are administered by<br />

<strong>the</strong> FS or BLM, but a fair portion is privately<br />

owned. Many of <strong>the</strong>se private lands are used <strong>for</strong><br />

cattle, which graze both in upland and adjoining<br />

riparian communities. Grazing in riparian<br />

communities is a concern because of <strong>the</strong> potential<br />

<strong>for</strong> negative impacts on areas that can provide<br />

dispersal habitat among mountain ranges.<br />

This mosaic pattern of jurisdiction by multiple<br />

landholders may impede coordinated management<br />

ef<strong>for</strong>ts <strong>for</strong> <strong>the</strong> owl.<br />

Most land development within <strong>the</strong> RU is<br />

related to enhancing recreation opportunities.<br />

These include developing or expanding campgrounds<br />

(e.g., Twilight Campground in <strong>the</strong><br />

Pinaleno Mountains, <strong>the</strong> loop turn at John<br />

Hand Lake in <strong>the</strong> Chiricahuas) or enlarging<br />

roads <strong>for</strong> safety (e.g., widening of Mount<br />

Lemmon highway in <strong>the</strong> Santa Catalina Mountains).<br />

Developments such as <strong>the</strong>se often require<br />

removing trees, potentially altering owl habitat.<br />

Fur<strong>the</strong>r, a rapidly increasing human population<br />

in <strong>the</strong> southwest portends increasing urban<br />

development which can potentially encroach<br />

upon owl habitat and also impact groundwater<br />

regimes, potentially impacting riparian systems.<br />

Evergreen oak and pinyon-juniper woodlands<br />

receive <strong>the</strong> most pressure from fuelwood<br />

harvest. Historical harvest of mature mesquite<br />

stands within <strong>the</strong> shrubland and grassland areas<br />

may have contributed to <strong>the</strong> demise of <strong>the</strong><br />

riparian <strong>for</strong>est, thus continued harvest of mature<br />

mesquite in <strong>the</strong>se areas may be a concern.<br />

Basin Basin Basin and and and Range Range Range - - - East East<br />

East<br />

The Basin and Range - East RU lies mostly<br />

within New Mexico and supports <strong>the</strong> second<br />

largest known number of <strong>Mexican</strong> spotted owls<br />

in <strong>the</strong> United States. It adjoins four U.S. and<br />

one <strong>Mexican</strong> RUs. <strong>Mexican</strong> spotted owls occur-<br />

Volume I/Part III<br />

102<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

ring in <strong>the</strong> Sacramento Mountains have been<br />

exposed to various disturbances <strong>for</strong> more than a<br />

century. Natural disturbances include <strong>for</strong>est fires<br />

plus insect and disease outbreaks. Human<br />

disturbances include timber and fuelwood<br />

harvest, grazing, land development, and recreation.<br />

The cumulative effects of <strong>the</strong>se natural<br />

and anthropogenic disturbances have resulted in<br />

a landscape that differs from that existing prior<br />

to European settlement. The threat of <strong>the</strong>se<br />

disturbances to <strong>the</strong> owl’s persistence cannot be<br />

quantified at this time, but certain incongruities<br />

are detectable. For example, owl density is<br />

relatively high on FS lands, but fecundity is<br />

quite variable over time and annual survival is<br />

unknown. Thus, even though <strong>the</strong> current<br />

population density may be high, we know<br />

nothing of population trends. Fur<strong>the</strong>r, given<br />

existing <strong>for</strong>est conditions in this RU, threats of<br />

widescale habitat loss are real and immediate.<br />

Consequently, active management is needed to<br />

alleviate <strong>the</strong>se threats while ensuring that adequate<br />

habitat will exist well into <strong>the</strong> future.<br />

Potential Potential Threats<br />

Threats<br />

The Team categorized potential threats to<br />

spotted owl recovery according to magnitude.<br />

Major threats pose immediate potential <strong>for</strong><br />

causing declines in spotted owl populations, and<br />

minor threats present no such immediacy. Major<br />

threats, in order of potential effects, include (1)<br />

catastrophic, stand-replacement fires, (2) some<br />

<strong>for</strong>ms of timber harvest, (3) fuelwood harvest,<br />

(4) grazing, (5) agriculture or development <strong>for</strong><br />

human habitation, and (6) <strong>for</strong>est insects and<br />

disease. Minor threats are activities not currently<br />

extensive in time or space but that have been<br />

considered potential threats to <strong>the</strong> owl. These<br />

include (1) certain military operations, (2) o<strong>the</strong>r<br />

habitat alterations (e.g. power line and road<br />

construction, noxious weed control), (3) mining,<br />

and (4) recreation.<br />

Existing dense <strong>for</strong>est conditions makes much<br />

of <strong>the</strong> Basin and Range - East RU vulnerable to<br />

catastrophic fire. Such stand-replacing fires have<br />

been documented in <strong>the</strong> Sacramento Mountains<br />

since <strong>the</strong> 1950s and continue to <strong>the</strong> present<br />

(e.g., <strong>the</strong> Burgett and Bridge fires in 1993 and<br />

1994, respectively). Similar fires occurred in <strong>the</strong>


Smokey Bear Ranger District (Hanks and Dick-<br />

Peddie 1974) and o<strong>the</strong>r mountain ranges in <strong>the</strong><br />

Basin and Range - East RU (Plummer and<br />

Gowsell 1904, Moody et al. 1992). Failure to<br />

address this potential <strong>for</strong> fire by reducing fuel<br />

levels and fuel continuity will inevitably lead to<br />

more and larger fires resulting in <strong>the</strong> continued<br />

loss of owl habitat.<br />

Past timber harvest practices have left a few<br />

remnant old-growth stands and residual pockets<br />

of pre-harvest trees in <strong>the</strong> Sacramento Mountains.<br />

Trees older than 200 ybh (years at breast<br />

height) can be found in <strong>the</strong>se remnant stands<br />

and pockets. Many of <strong>the</strong>se stands, however, are<br />

small (


indicates that spotted owls use <strong>for</strong>est types not<br />

typically found in <strong>the</strong> United States and that<br />

land-management practices differ substantially<br />

in Mexico from those in <strong>the</strong> U.S. The Team<br />

proposes that <strong>the</strong> general recommendations be<br />

applied within <strong>the</strong> <strong>Mexican</strong> RUs where appropriate.<br />

The Team strongly recommends that RU<br />

working teams (see Part IV) develop management<br />

recommendations <strong>for</strong> Mexico more fully.<br />

The <strong>Mexican</strong> spotted owl is typically found<br />

in montane habitats where <strong>the</strong> vegetation is<br />

dominated by pines and oaks. Although spotted<br />

owls in <strong>the</strong> U.S. also use pine-oak <strong>for</strong>ests, <strong>the</strong>y<br />

vary somewhat in composition and structure<br />

from similar <strong>for</strong>ests in Mexico. However, within<br />

<strong>Mexican</strong> pine-oak <strong>for</strong>ests, spotted owls appear to<br />

favor canyons, as <strong>the</strong>y do in many of <strong>the</strong> areas<br />

used in <strong>the</strong> U.S.<br />

Because social and economic systems in<br />

Mexico differ from those in <strong>the</strong> U.S., activities<br />

that take place within potential spotted owl<br />

habitat differ somewhat between <strong>the</strong> two countries.<br />

Whereas grazing, fire, timber harvest, and<br />

fuelwood harvest are threats common to both<br />

countries, o<strong>the</strong>r threats are unique to Mexico<br />

and some to specific RUs.<br />

Potential Potential Threats<br />

Threats<br />

Sierra Sierra Madre Madre Occidental Occidental - - Norte Norte.— Norte .— .—The .— primary<br />

threat is land conversion <strong>for</strong> subsistence<br />

agriculture. Impacts of this threat include <strong>the</strong><br />

loss of spotted owl habitat, soil loss, and erosion.<br />

A related threat is overgrazing by livestock.<br />

Historically and through <strong>the</strong> present, <strong>for</strong>est<br />

lands have been cleared to create pastures <strong>for</strong><br />

cattle; <strong>the</strong> cumulative effects of <strong>the</strong>se practices<br />

have modified spotted owl habitat. Although<br />

extensive timber harvest occurs within this RU,<br />

most harvest occurs in <strong>the</strong> uplands and is not a<br />

direct threat to spotted owls found in canyons.<br />

Whe<strong>the</strong>r or not timber harvest indirectly affects<br />

<strong>the</strong> owl is unknown, but could be a concern<br />

worth addressing through research.<br />

Sierra Sierra Sierra Madre Madre Oriental Oriental - - Norte Norte.— Norte .— .—Timber .—<br />

harvest and grazing are <strong>the</strong> primary threats<br />

within this RU. The main effects of <strong>the</strong>se threats<br />

is to fur<strong>the</strong>r fragment an already disjunct population.<br />

Volume I/Part III<br />

104<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Sierra Sierra Madre Madre Occidental Occidental - - Sur Sur.— Sur .— .—Primary .—<br />

threats within this RU are fuelwood harvest,<br />

timber harvest, charcoal production, grazing,<br />

and agricultural development. Fuelwood harvest<br />

often entails cutting snags and <strong>the</strong> harvest of<br />

riparian plant species, both of which are critical<br />

components of spotted owl habitat. Timber<br />

harvest in itself is not a direct threat to spotted<br />

owl habitat. However, harvest methods that<br />

entail rolling logs from higher to lower sites<br />

result in soil loss and erosion, <strong>the</strong>reby affecting<br />

<strong>the</strong> habitat of <strong>the</strong> owl and its prey. Fire is considered<br />

only a moderate threat because of <strong>the</strong><br />

disjunct distribution of owls in canyons and<br />

because limited ef<strong>for</strong>ts towards fire suppression<br />

have allowed natural fire regimes to persist. Fire,<br />

however, can possibly destroy spotted owl<br />

habitat under <strong>the</strong> right conditions. Livestock<br />

graze throughout <strong>the</strong> year within spotted owl<br />

habitat, and <strong>the</strong> cumulative effect of this grazing<br />

affects prey habitat and spotted owl habitat<br />

structure. Type conversion of <strong>for</strong>ests <strong>for</strong> agriculture<br />

also occurs within this RU.<br />

Sierra Sierra Madre Madre Oriental Oriental Oriental - - - Sur Sur.— Sur .— .—The .— potential<br />

threats in this RU parallel those within Sierra<br />

Madre Occidental - Norte RU. The primary<br />

difference is that this RU probably contains <strong>the</strong><br />

best and most extensive habitat within any of <strong>the</strong><br />

<strong>Mexican</strong> RUs, thus threats can occur over a<br />

much greater area.<br />

Eje Eje Neovolcanico<br />

Neovolcanico.—<br />

Neovolcanico .— .—The .— primary threats in this<br />

RU are those associated with population expansion<br />

and industrial development. Specifically,<br />

industrial development can lead to loss of habitat<br />

and an increase in pollution. O<strong>the</strong>r threats<br />

include management to control insects and<br />

disease, fire suppression, grazing, and agricultural<br />

development.


The Team has assimilated, reviewed, and<br />

analyzed data generated by <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong><br />

<strong>Owl</strong> Monitoring Program of FS Region 3. We<br />

have also compiled and reviewed data from <strong>the</strong><br />

BLM and <strong>the</strong> FS Region 4 in Utah, and FS<br />

Region 2 in Colorado. Here, we offer an alternative<br />

design <strong>for</strong> monitoring <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl population within <strong>the</strong> three core RUs. We<br />

also provide recommendations <strong>for</strong> monitoring<br />

habitat throughout <strong>the</strong> owl’s range. This proposed<br />

monitoring program will evaluate population<br />

and habitat trends as required by <strong>the</strong> criteria<br />

<strong>for</strong> delisting <strong>the</strong> species (III.A). The philosophy<br />

of our proposed monitoring scheme is to measure<br />

<strong>the</strong> critical variables--changes in owl numbers<br />

and changes in habitat--needed <strong>for</strong> delisting<br />

<strong>the</strong> species.<br />

For <strong>the</strong> purposes of recovery and understanding<br />

effects of land management activities<br />

on <strong>the</strong> <strong>Mexican</strong> spotted owl, monitoring should<br />

determine with adequate reliability temporal<br />

changes in <strong>the</strong> owl population and its habitat<br />

when in fact such changes are occurring. An<br />

effective monitoring program requires measuring<br />

changes in habitat quantity, estimating population<br />

size of territorial owls, and determining key<br />

demographic parameters including survival,<br />

recruitment, and reproduction, all of which<br />

influence population size.<br />

Habitat monitoring will rely heavily on both<br />

remote-sensing of habitat across <strong>the</strong> range of <strong>the</strong><br />

bird and field measurements of habitat variables<br />

be<strong>for</strong>e and after treatments. Population monitoring<br />

is based on mark-recapture <strong>the</strong>ory and<br />

Cormack-Jolly-Seber modeling approaches,<br />

similar to Pollock’s robust design approach<br />

(Lefebre et al. 1982, Pollock 1982, Kendall and<br />

Pollock 1992). To our knowledge, a systematic<br />

habitat monitoring approach as extensive and<br />

intensive as <strong>the</strong> one we propose has never been<br />

implemented. In contrast, a prototype design <strong>for</strong><br />

population monitoring was implemented in<br />

Olympic National Park, Washington (Noon et<br />

al. 1993, E. Seamen pers. comm.) to monitor a<br />

nor<strong>the</strong>rn spotted owl population.<br />

Volume I/Part III<br />

C. C. MONITORING MONITORING PROCEDURES<br />

PROCEDURES<br />

105<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Accurate and efficient protocols <strong>for</strong> both<br />

habitat and population monitoring require pilot<br />

studies to estimate recapture probabilities, and to<br />

estimate variances associated with each of <strong>the</strong><br />

population parameters and each of <strong>the</strong> habitat<br />

variables. For population monitoring, <strong>the</strong>se<br />

estimates can <strong>the</strong>n be used to determine optimal<br />

quadrat size and numbers of quadrats required<br />

within predefined strata and RUs. Funding was<br />

allocated in 1994 to fur<strong>the</strong>r refine <strong>the</strong> proposed<br />

population design with a small field trial involving<br />

four quadrats. Results of this field trial<br />

validated <strong>the</strong> study design provided qualified<br />

personnel conduct <strong>the</strong> work (May et al., in<br />

press). A larger pilot study is needed to refine<br />

parameter estimates <strong>for</strong> actual implementation<br />

of <strong>the</strong> proposed design. For habitat monitoring,<br />

separate pilot studies will be needed to establish<br />

sampling designs, including sample size requirements.<br />

HABITAT HABITAT MONITORING<br />

MONITORING<br />

The habitat delisting criterion states that<br />

habitat monitoring must be implemented (1) to<br />

track changes in <strong>the</strong> quantity of macrohabitat<br />

and (2) to verify that microhabitat changes<br />

within treated stands meet <strong>the</strong> intent of <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong>. Thus little, if any, owl habitat can<br />

be lost if this goal is to be met. Fur<strong>the</strong>r, habitat<br />

quality cannot decline significantly. A concern of<br />

<strong>the</strong> Team is that habitat quality cannot be<br />

adequately assessed, particularly with remotesensing<br />

data. To alleviate this concern, our<br />

recommendations also include field measurements<br />

of microhabitat characteristics within<br />

treated stands. However, we reiterate that<br />

macrohabitat quantity should also be monitored<br />

on a rangewide basis.<br />

Macrohabitat<br />

Macrohabitat<br />

Macrohabitat<br />

The purpose of rangewide monitoring is to<br />

track gross changes in habitat as <strong>the</strong> result of<br />

disturbance, from both natural (e.g., fire) and


anthropogenic (e.g., timber harvest, prescribed<br />

fire) causes. Given <strong>the</strong> extent of <strong>the</strong> area to be<br />

monitored, remote-sensing technology will be<br />

required. An imagery baseline should be established<br />

within six months of <strong>Recovery</strong> <strong>Plan</strong><br />

approval using LANDSAT Thematic Mapper<br />

imagery (currently available in <strong>the</strong> FS Region 3<br />

Geometronics Remote-Sensing Laboratory). The<br />

imagery <strong>for</strong> potential owl habitat and surrounding<br />

areas should be aggregated, georeferenced,<br />

and merged with vector map data to provide<br />

image maps. These image maps may be used<br />

both as general planning tools and as a baseline<br />

<strong>for</strong> detecting change. The 30-m (98-ft) spatial<br />

resolution of <strong>the</strong> LANDSAT Thematic Mapper<br />

imagery is adequate to detect both anthropogenic<br />

and natural change across large land areas.<br />

Changes that can be detected include wildfire<br />

scars, timber harvests, gross changes in <strong>for</strong>est<br />

health, and possibly <strong>the</strong> addition or removal of<br />

roads, large developments and o<strong>the</strong>r cultural<br />

features, and fluctuations in grazing practices.<br />

In a standard change-detection analysis, two<br />

image data sets are co-registered and <strong>the</strong>n subtracted<br />

from each o<strong>the</strong>r. The resulting difference<br />

image highlights changes across <strong>the</strong> area covered<br />

by <strong>the</strong> image data sets. An additional image data<br />

set will need to be purchased in <strong>the</strong> future to<br />

per<strong>for</strong>m change detection. The additional data<br />

set may be ei<strong>the</strong>r LANDSAT Thematic Mapper<br />

(TM) data or LANDSAT Multispectral Scanner<br />

(MSS) data. A TM data set would be preferred,<br />

but MSS data could be substituted and would<br />

provide adequate spectral and spatial resolution<br />

to per<strong>for</strong>m useful change detection.<br />

The baseline TM data set can also be used to<br />

develop a generalized regional vegetation cover<br />

map, using standard supervised and unsupervised<br />

image classification techniques with TM<br />

bands 4, 3, and 2. The literature indicates that<br />

<strong>the</strong> 4,3,2 band combination is most useful <strong>for</strong><br />

vegetation analysis. Developing a vegetation map<br />

would also require <strong>the</strong> integration of 1:250,000<br />

digital elevation data to account <strong>for</strong> <strong>the</strong> effects of<br />

varying slope aspects and elevation on vegetation<br />

patterns.<br />

Texture analysis techniques may be used to<br />

assess vegetation structure and density across<br />

Volume I/Part III<br />

106<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

large areas using ei<strong>the</strong>r <strong>the</strong> LANDSAT TM or<br />

MSS data. While texture analysis applications in<br />

vegetation analysis have appeared frequently in<br />

<strong>the</strong> literature, <strong>the</strong> methodologies are not as<br />

widely accepted as image classification techniques,<br />

so <strong>the</strong>y must be considered experimental.<br />

Given <strong>the</strong> resolution possible with <strong>the</strong> tools<br />

and in<strong>for</strong>mation available, remote sensing<br />

monitoring techniques will provide an estimate<br />

of macrohabitat trend. Within five years of<br />

creating <strong>the</strong> imagery baseline, participating<br />

agencies should produce a report assessing<br />

changes in vegetation composition, structure,<br />

and density. This will provide an interim checkpoint<br />

to determine if <strong>the</strong> delisting criteria can be<br />

met at <strong>the</strong> end of 10 years.<br />

Microhabitat<br />

Microhabitat<br />

Microhabitat monitoring is required because<br />

remote sensing is largely insensitive to subtle<br />

intra-stand changes that may enhance or degrade<br />

owl habitat. Microhabitat monitoring will entail<br />

measuring habitat variables be<strong>for</strong>e and after<br />

silvicultural or prescribed fire treatments designed<br />

to maintain, improve, or create owl<br />

habitat. This monitoring is to verify that treatments<br />

(silviculture, fire) are meeting <strong>the</strong>ir stated<br />

objectives. We acknowledge that many treated<br />

stands will not meet <strong>the</strong> desired future condition<br />

in 10 years, but <strong>the</strong> trajectory on which a stand<br />

is placed can be modeled to evaluate if it is<br />

moving towards owl habitat. This knowledge is<br />

needed to demonstrate that any short-term losses<br />

in macrohabitat will be partially offset as stands<br />

mature into owl habitat. If adequate acreage of<br />

vegetation is moving towards owl habitat, our<br />

confidence in long-term habitat stability will<br />

be enhanced.<br />

Sampling units will be treated stands. Within<br />

<strong>the</strong>se stands, an adequate number of vegetation<br />

sampling points must be established. The exact<br />

number of sampling points needed will be<br />

dictated by <strong>the</strong> most variable characteristic<br />

(likely snag density; Bull et al. 1990). Points<br />

should be sampled prior to initiating a treatment,<br />

resampled following <strong>the</strong> treatment after<br />

allowing adequate time <strong>for</strong> <strong>the</strong> area to equilibrate


from temporary disturbance effects, and <strong>the</strong>n at<br />

five-year intervals. The variables measured can<br />

<strong>the</strong>n be input into a vegetation model to estimate<br />

stand characteristics at different points in<br />

time.<br />

At a minimum <strong>the</strong> following variables<br />

should be measured and assessed: (1) tree diameters<br />

by species, (2) tree basal area by species, (3)<br />

size-class distributions of trees, (4) log volume by<br />

size class, (5) canopy cover, (6) snag diameter,<br />

and (7) snag basal area. We strongly advocate<br />

that additional variables be included that might<br />

be relevant to monitoring o<strong>the</strong>r ecosystem<br />

attributes. The return in critical monitoring<br />

in<strong>for</strong>mation derived by expanding <strong>the</strong> variables<br />

measured would far outweigh any additional<br />

costs, assuming that <strong>the</strong> new variables are not<br />

highly correlated with <strong>the</strong> variables suggested<br />

above.<br />

POPULATION POPULATION MONITORING<br />

MONITORING<br />

Monitoring habitat as a singular ef<strong>for</strong>t will<br />

not adequately reveal <strong>the</strong> true status of <strong>the</strong> owl<br />

population. Relatively long-lived birds with a<br />

high (~0.89) adult survival, <strong>Mexican</strong> spotted<br />

owls may live 16 years or more once <strong>the</strong>y reach<br />

adulthood. However, an intense period of<br />

mortality during <strong>the</strong> first year could produce<br />

population consequences that habitat monitoring<br />

would not detect. Habitat quality could<br />

decline from various natural processes or anthropogenic<br />

activities, yet <strong>the</strong> territorial population<br />

would remain unchanged because of site fidelity<br />

among existing birds and recruitment of floaters.<br />

Young might still be produced, but would not<br />

survive to be recruited into <strong>the</strong> territorial population<br />

because of poor habitat quality, limited<br />

habitat availability, or because <strong>the</strong>ir inexperience<br />

would not allow <strong>the</strong>m to survive and disperse<br />

during <strong>the</strong>ir first year.<br />

A limitation of this proposed monitoring<br />

scheme (and all known approaches) is that only<br />

territorial birds are monitored. The total or<br />

proportional number of floaters (sensu Franklin<br />

1992) remains undetermined and unmonitored<br />

relative to <strong>the</strong> target population.<br />

Volume I/Part III<br />

107<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Because nonterritorial birds are not directly<br />

monitored, we want to guard against an undetected<br />

decline in <strong>the</strong> total population of spotted<br />

owls when <strong>the</strong> territorial population remains<br />

stable. We suggest <strong>the</strong> following procedures to<br />

evaluate trends in <strong>the</strong> nonterritorial population.<br />

First, <strong>the</strong> age of birds that establish territories<br />

will indicate <strong>the</strong> size of <strong>the</strong> nonterritorial population.<br />

If new territorial birds are only one year<br />

old, <strong>the</strong>n <strong>the</strong>y have never existed as floaters in<br />

<strong>the</strong> population. Thus, a decline in <strong>the</strong> age of<br />

territorial birds suggests that <strong>the</strong> nonterritorial<br />

population is low or declining (Franklin 1992).<br />

Second, <strong>the</strong> presence of unfilled territories would<br />

suggest that an inadequate floater population<br />

exists, and hence that a decline in <strong>the</strong> population<br />

is taking place.<br />

In <strong>the</strong> following, we outline a suitable<br />

framework and statistical estimation approach<br />

<strong>for</strong> monitoring owl populations in 3 RUs.<br />

However, critical design and sampling details,<br />

such as sample sizes and delineation of strata,<br />

have been omitted, and must be developed by an<br />

implementation team as data become available<br />

to make those decisions.<br />

Target Target Population<br />

Population<br />

The target population <strong>for</strong> <strong>the</strong> abundance<br />

estimate is territorial <strong>Mexican</strong> spotted owls<br />

(exclusive of floaters) in <strong>the</strong> Upper Gila Mountains,<br />

Basin and Range - West, and Basin and<br />

Range - East RUs. Thus, all potential owl habitat<br />

in <strong>the</strong>se 3 RUs must be included in <strong>the</strong> sampling<br />

frame. All land management jurisdictions are<br />

encouraged to cooperate in providing access and<br />

resources to monitor <strong>the</strong> entire owl population.<br />

Sampling Sampling Units<br />

Units<br />

Sampling units will consist of 50 to 75 km 2<br />

(19 to 29 mi 2 ) quadrats randomly allocated to<br />

habitat strata. Quadrats will be defined based on<br />

ecological boundaries such as ridge lines and<br />

watersheds to reduce edge effects. Selection of<br />

quadrat boundaries must emphasize edges that<br />

are unlikely to traverse owl territories, so that <strong>the</strong><br />

errors of including a territory in multiple quad-


ats or in no quadrat do not occur. The exact<br />

number of quadrats and <strong>the</strong>ir size will depend<br />

on <strong>the</strong> specifics of implementing <strong>the</strong> monitoring<br />

scheme and results of pilot studies.<br />

In general, <strong>the</strong> population monitoring<br />

scheme will require: (1) determining strata that<br />

represent different owl densities or habitat types<br />

occupied by owls within each RU; (2) determining<br />

quadrat size which should be sufficiently<br />

large to reduce edge effects and small enough to<br />

allow a minimum of four surveys per quadrat in<br />

<strong>the</strong> survey season limited to 1 April to 30 August<br />

(initial approximation of quadrat size is 50 to<br />

75 km 2 [19 to 29 mi 2 ] [May et al., in press]);<br />

(3) defining <strong>the</strong> sampling frame of quadrats in<br />

each stratum such that quadrats are relatively<br />

equal in size and have boundaries selected to<br />

minimize edge effect; (4) selecting a random<br />

sample of quadrats from each stratum <strong>the</strong> first<br />

year, and <strong>the</strong>n randomly replacing 20% of <strong>the</strong><br />

sampled quadrats each year with quadrats<br />

randomly selected from <strong>the</strong> currently unsampled<br />

quadrats (with quadrats in <strong>the</strong> initial sample<br />

potentially removed, and <strong>the</strong>n possibly included<br />

in <strong>the</strong> sample again at a later time); and (5)<br />

developing protocols <strong>for</strong> conducting field<br />

surveys (likely different from past FS protocols<br />

because of <strong>the</strong> different goal of <strong>the</strong> proposed<br />

procedure from past goals).<br />

Sampling Sampling Procedures<br />

Procedures<br />

Stratification<br />

Stratification<br />

LANDSAT multispectral scanner imagery<br />

with 30-m spatial resolution would be suitable<br />

<strong>for</strong> defining habitat strata within each RU, and<br />

thus <strong>the</strong> sampling frame of quadrats <strong>for</strong> each<br />

stratum. Approximate density of territorial owls<br />

within strata can <strong>the</strong>n be used to allocate survey<br />

ef<strong>for</strong>t (number of quadrats) to strata. The<br />

optimum allocation of survey ef<strong>for</strong>t is one that<br />

would minimize erroneous estimation of spotted<br />

owl abundance. Optimal allocation of quadrats<br />

to strata will probably not be in proportion to<br />

strata size. More likely, optimal allocation will<br />

mean that a higher percentage of quadrats in<br />

strata with high owl densities will be sampled.<br />

Volume I/Part III<br />

108<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Optimal allocation might also take into account<br />

<strong>the</strong> cost per quadrat because of potential differences<br />

in <strong>the</strong> cost of measuring quadrats within<br />

different strata. Strata should be computed both<br />

as projected (ignoring topography) and as<br />

surface (incorporating topography) areas because<br />

differences in topography affect vegetation type.<br />

Selection Selection of of Quadrats<br />

Quadrats<br />

Within strata, quadrats will be randomly<br />

selected <strong>for</strong> inclusion in <strong>the</strong> sample. We suggest<br />

that 80% (randomly selected) of <strong>the</strong> previous<br />

year’s quadrats be revisited <strong>the</strong> following year.<br />

The o<strong>the</strong>r 20% of <strong>the</strong> previous year’s quadrats<br />

should be removed from <strong>the</strong> sample, and a<br />

replacement sample of quadrats not sampled <strong>the</strong><br />

previous year should be substituted. Quadrats<br />

removed one year can included in <strong>the</strong> sample in<br />

following years provided that it is randomly<br />

selected and that at least one year has elapsed<br />

since it was last sampled. This procedure means<br />

that <strong>the</strong> average number of years that a particular<br />

quadrat remains in <strong>the</strong> sample is 4.5 years.<br />

Inclusion of new quadrats into <strong>the</strong> sample each<br />

year guards against management practices within<br />

<strong>the</strong> sampled quadrats not being representative of<br />

those practices occurring elsewhere. This rotation<br />

of quadrats included in <strong>the</strong> sample will<br />

provide a smaller variance because observations<br />

are correlated across time and many quadrats will<br />

be sampled repeatedly, but still provides a<br />

representative sample of <strong>the</strong> available quadrats.<br />

Sampling Sampling Within Within Quadrats<br />

Quadrats<br />

Sampling procedures within a quadrat will<br />

include (1) assigning survey stations to ensure<br />

adequate coverage and a standardized density of<br />

such stations among quadrats; (2) allowing <strong>for</strong> a<br />

minimum of four complete (i.e., all call points<br />

sampled) surveys through each quadrat; (3)<br />

conducting nighttime surveys from survey<br />

stations to map general locations of spotted owls<br />

and to estimate per-visit detection probabilities;<br />

and (4) conducting daytime (auxiliary) surveys<br />

and mousing to find roosting and nesting<br />

spotted owls, determine if a mate not detected


during nighttime survey is present, determine<br />

number of young present, and capture and<br />

color-mark all spotted owls found. For each<br />

spotted owl found, <strong>the</strong> center of its activity area<br />

must be determined as ei<strong>the</strong>r in or out of <strong>the</strong><br />

quadrat.<br />

Proper timing of surveys maximizes efficiency<br />

in locating spotted owls. We recommend<br />

that nighttime surveys be conducted within 3<br />

hours following sunset. <strong>Owl</strong>s detected at dusk<br />

are near diurnal roosts and thus provide an<br />

optimal starting point <strong>for</strong> confirming pairs and<br />

reproduction during daytime surveys. Similarly,<br />

daytime surveys should begin at or near sunrise<br />

(preferably just be<strong>for</strong>e) as owls are returning to<br />

<strong>the</strong>ir roosts.<br />

Banding Banding Banding Birds<br />

Birds<br />

Marking individual birds with FWS leg<br />

bands and color bands <strong>for</strong> visual identification<br />

provides greater validity in <strong>the</strong> estimation of <strong>the</strong><br />

owl population size on <strong>the</strong> quadrat because <strong>the</strong><br />

assumptions of <strong>the</strong> mark-recapture methods can<br />

be tested. Conceivably, owl population size on<br />

quadrats with high densities of owls might be<br />

underestimated without banding because two<br />

different birds might be counted as only one.<br />

Conversely, on quadrats with low densities, a<br />

single bird might be counted as 2 birds, biasing<br />

<strong>the</strong> population estimate high. Individually<br />

marking birds will eliminate some of this potential<br />

bias. Second, banding birds is necessary to<br />

estimate annual survival on quadrats that are<br />

sampled <strong>for</strong> two consecutive years. Third,<br />

capturing birds allows more careful aging of<br />

individuals; hence, <strong>the</strong> resulting age structure<br />

data are more useful in assessing <strong>the</strong> impact of<br />

floaters in <strong>the</strong> population. Finally, minimum<br />

estimates of dispersal and emigration from <strong>the</strong><br />

quadrat can be assessed with banded birds that<br />

are located off <strong>the</strong> quadrat. Although <strong>the</strong> cost of<br />

<strong>the</strong> population estimation procedure may be<br />

increased by up to 40% by individually marking<br />

birds (May et al., in press), <strong>the</strong> Team feels <strong>the</strong><br />

additional in<strong>for</strong>mation and rigor provided by<br />

marking birds is justified.<br />

Volume I/Part III<br />

109<br />

Statistical Statistical Analysis<br />

Analysis<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Noon et al. (1993) provide <strong>the</strong> following<br />

statistical discussion.<br />

Estimation Estimation of of of Capture Capture Probability Probability<br />

Probability<br />

The described spotted owl surveys will<br />

provide data regarding <strong>the</strong> number of territorial<br />

owls detected and determined to have <strong>the</strong>ir<br />

activity centers within <strong>the</strong> survey plots. It is<br />

unlikely that all owls with activity foci lying<br />

within <strong>the</strong>se plots will be detected, and capture<br />

probabilities will <strong>the</strong>re<strong>for</strong>e be less than 1. Capture<br />

is defined as ei<strong>the</strong>r physically capturing and<br />

uniquely marking an individual or resighting its<br />

unique color band combination without physically<br />

recapturing it. Thus, <strong>the</strong> per-visit capture<br />

probabilities must be estimated to “correct” <strong>the</strong><br />

count statistics and reflect <strong>the</strong> true number of<br />

territorial birds with activity centers within<br />

quadrat boundaries.<br />

Per-visit capture probabilities can be estimated<br />

using data on <strong>the</strong> capture histories of<br />

individual owls on <strong>the</strong> quadrats. The four<br />

surveys will be conducted during a relatively<br />

short period, so it is appropriate to use capturerecapture<br />

models <strong>for</strong> closed populations (e.g.,<br />

Otis et al. 1978, White et al. 1982, Pollock et al.<br />

1990). Per-visit capture probabilities may vary by<br />

visits, strata, RUs, and years. However, substantial<br />

gains in <strong>the</strong> precision of capture probability<br />

estimates would be achieved if <strong>the</strong>y could be<br />

estimated using data pooled over visits, strata,<br />

RUs, or years. Standardized survey protocol<br />

within and among quadrats using field crews<br />

with communal training should decrease <strong>the</strong><br />

variation in per-visit capture probabilities.<br />

Heterogeneity of per-visit capture probabilities<br />

across individuals should be examined. If<br />

heterogeneity is found, estimators developed<br />

under model M h of Otis et al. (1978), such as<br />

<strong>the</strong> jackknife (Burnham and Overton 1978,<br />

1979) and Chao’s (1987, 1988, 1989) estimator<br />

are appropriate. If heterogeneity is not serious,<br />

models with data pooled over visits, strata, RUs,<br />

and years should be considered.<br />

Capture probability estimates resulting from<br />

<strong>the</strong>se modeling ef<strong>for</strong>ts will pertain to <strong>the</strong> prob


ability of detecting and capturing an individual<br />

spotted owl during a single survey visit. Capture<br />

history data, and hence <strong>the</strong> capture probability<br />

estimates, will be restricted to those spotted owls<br />

with a nest or focus of activity within <strong>the</strong> area<br />

being sampled. The density estimation procedure<br />

(see below) actually requires an estimate of<br />

<strong>the</strong> probability of detecting and capturing a<br />

spotted owl at least once during an entire season.<br />

Given owl presence, <strong>the</strong> estimated probability of<br />

detecting and capturing an owl during <strong>the</strong><br />

season using surveys is given by<br />

Volume I/Part III<br />

p = 1 - (1 - p) k k ^<br />

,<br />

where p k is <strong>the</strong> probability of detecting and<br />

capturing a spotted owl at least once during<br />

k visits, and p is <strong>the</strong> single-visit capture probability.<br />

The above scheme <strong>for</strong> estimating capture<br />

probability should work well with owls initially<br />

detected from survey points within <strong>the</strong> quadrat.<br />

However, <strong>the</strong>se capture probabilities are not<br />

applicable, by <strong>the</strong>mselves, to o<strong>the</strong>r pair members<br />

found during daytime visits or to spotted owls<br />

located o<strong>the</strong>r than by calling from survey points.<br />

To include data from pair members located<br />

during daytime visits, we consider <strong>the</strong>ir capture<br />

probability as <strong>the</strong> product of <strong>the</strong> probability of<br />

capturing a pair member from survey points<br />

times <strong>the</strong> probability of capturing <strong>the</strong> o<strong>the</strong>r pair<br />

member during a daytime visit given capture of<br />

one mate from survey points. The first probability<br />

in this product is obtained using capture<br />

histories (i.e., capture probabilities) of birds<br />

detected from survey points as previously described.<br />

The second, conditional probability<br />

must be obtained using data from daytime visits<br />

and captures only.<br />

Auxiliary or daytime visits can be viewed in<br />

<strong>the</strong> context of removal modeling (e.g., Zippin<br />

1956, 1958; Otis et al. 1978; Ward et al. 1991).<br />

The primary purpose of daytime visits is to<br />

determine pair and breeding status of birds<br />

detected from survey points. Auxiliary visits may<br />

or may not be terminated after capture of <strong>the</strong><br />

mate, depending on whe<strong>the</strong>r data on reproductive<br />

success is adequately obtained.<br />

110<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Estimation Estimation Estimation of of of Density Density Per Per Quadrat<br />

Quadrat<br />

Each quadrat requires two parts to <strong>the</strong><br />

estimation process. First, estimation of apparent<br />

density by<br />

^<br />

Di n n· + n¨ a i i i<br />

= = ,<br />

which is <strong>the</strong> total number of spotted owls<br />

detected and having activity centers in quadrat i<br />

·<br />

¨<br />

(n i ), based on both survey (n i ) and auxiliary (n i )<br />

visits divided by <strong>the</strong> area of <strong>the</strong> quadrat (a i ).<br />

Next, apparent density is adjusted <strong>for</strong> those<br />

spotted owls that maintain an activity center<br />

within <strong>the</strong> quadrat but were not detected during<br />

a survey visit. The adjustment <strong>for</strong> <strong>the</strong> survey visit<br />

is <strong>the</strong> reciprocal of p , <strong>the</strong> probability of a<br />

k<br />

spotted owl being captured at least once on a<br />

given survey based on k visits to quadrat i. Note<br />

that p pertains only to <strong>the</strong> n animals detected<br />

k i<br />

during <strong>the</strong> survey point visits of quadrat i, not to<br />

birds detected on auxiliary visits. If auxiliary<br />

visits are made to determine pair or reproductive<br />

status, any additional pair members that are<br />

detected contribute to <strong>the</strong> density estimate <strong>for</strong><br />

quadrat i. As discussed previously, <strong>the</strong>ir capture<br />

is conditional upon an initial capture from a<br />

survey point. To adjust <strong>the</strong> count of <strong>the</strong> n owls i<br />

detected on quadrat i during auxiliary visits, we<br />

divide <strong>the</strong> count by <strong>the</strong> probability of detecting a<br />

pair member during k auxiliary visits (p ), given<br />

k<br />

detection and capture of its mate from survey<br />

points. Thus, we would estimate density on<br />

quadrat i as<br />

ni ^ p^ D = k .<br />

i + ni ¨<br />

^<br />

^<br />

·<br />

¨<br />

¨<br />

·<br />

p^ p¨ k k<br />

Estimation Estimation of of Density Density Per Per Stratum<br />

Stratum<br />

Once we have density estimates <strong>for</strong> each<br />

quadrat within a stratum, <strong>the</strong>y can be combined<br />

into an overall mean estimate of density <strong>for</strong> <strong>the</strong><br />

stratum. Because quadrats are not <strong>the</strong> same size,<br />

weighting of <strong>the</strong> quadrat density estimates by<br />

area is essential, so that<br />

D j =<br />

m j<br />

S<br />

a i<br />

a i<br />

^<br />

a D i=1 i i ,<br />

mj Sai<br />

i=1<br />

a i


where m is <strong>the</strong> number of quadrats in stratum j.<br />

j<br />

To obtain population size estimates <strong>for</strong> each<br />

stratum (j), multiply by <strong>the</strong> area of <strong>the</strong> stratum<br />

^<br />

(A ) so that N = D · A . Note that A may change<br />

j j j j j<br />

through time as habitat changes and quadrats are<br />

moved from one stratum to ano<strong>the</strong>r.<br />

An estimate of overall abundance <strong>for</strong> RU u is<br />

mu ^ ^<br />

<strong>the</strong>n N = S N , where m is <strong>the</strong> number of<br />

u j=1 j u<br />

strata in RU u.<br />

Variance Variance Estimators<br />

Estimators<br />

With stratified-random sampling, we usually<br />

have simple variance equations because <strong>the</strong><br />

stratum means of totals are independent between<br />

strata. Here, this is not <strong>the</strong> case because corrections<br />

<strong>for</strong> spotted owls not seen are common<br />

across strata and induce <strong>the</strong> need <strong>for</strong> covariance<br />

terms. These variance equations will need to be<br />

developed. Specific closed-<strong>for</strong>med solutions may<br />

not be possible <strong>for</strong> <strong>the</strong> variance estimates. Thus,<br />

estimating <strong>the</strong> variance components by bootstrap<br />

methods may be more feasible.<br />

Cormack-Jolly-Seber Cormack-Jolly-Seber Models<br />

Models<br />

Cormack-Jolly-Seber (CJS) models<br />

(Lebreton et al. 1992) can be used to estimate<br />

age-specific apparent survival (f a ) and recruitment<br />

to <strong>the</strong> territorial population (B). We<br />

suggest using CJS modeling procedures as<br />

outlined by Lebreton et al. (1992), Burnham<br />

and Anderson (1992), Pollock et al. (1990), and<br />

Burnham et al. (1987). This approach is demonstrated<br />

by White et al. (1995) <strong>for</strong> <strong>the</strong> analysis of<br />

data from <strong>the</strong> demographic study areas. We<br />

suggest that data from all quadrats within a<br />

stratum (or even larger area) can be pooled to<br />

estimate apparent survival and recruitment.<br />

Both <strong>the</strong> apparent survival (f) and recruitment<br />

(B) are biased estimates of true survival (S)<br />

and true recruitment rates because <strong>the</strong> quadrats<br />

do not have geographic closure. For survival,<br />

f = S - E, where E is <strong>the</strong> emigration rate off <strong>the</strong><br />

quadrats. For recruitment, B = R + I, where R is<br />

true recruitment and I is immigration onto <strong>the</strong><br />

quadrats. If <strong>the</strong> area of <strong>the</strong> combined quadrats<br />

were used in a single demographic study area,<br />

<strong>the</strong> bias of f and B would be smaller because <strong>the</strong><br />

Volume I/Part III<br />

111<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

probability of birds emigrating off of and immigrating<br />

onto a large single area would be smaller<br />

than <strong>for</strong> a collection of small quadrats representing<br />

<strong>the</strong> same area. However, if we assume that<br />

this bias is somewhat constant across time, <strong>the</strong>n<br />

tests <strong>for</strong> changes in f and B across time with<br />

models such as f T as demonstrated by Burnham<br />

et al. (1994) provide a potent tool to assess<br />

changes in <strong>the</strong>se population parameters through<br />

time. The optimal size of quadrats is dictated by<br />

keeping <strong>the</strong>m large enough that reasonable<br />

estimates of <strong>the</strong> number of territorial birds<br />

present can be accomplished, while small<br />

enough so that an adequately large sample of<br />

quadrats is possible to estimate precisely <strong>the</strong><br />

among-quadrat variation.<br />

We would not suggest that f and B be used<br />

to compute l in a Leslie matrix model as was<br />

done with <strong>the</strong> demographic study areas by<br />

White et al. (1995). Biases caused by emigration<br />

and immigration make any estimate of l computed<br />

from <strong>the</strong>se parameters biased as well.<br />

Fur<strong>the</strong>rmore, <strong>the</strong> main objective of <strong>the</strong> quadrat<br />

surveys is to provide an unbiased estimate of <strong>the</strong><br />

total number of territorial owls so that an<br />

unbiased estimate of l can be obtained as<br />

^ ^ ^<br />

l = N t+1 /N t .<br />

Estimates of juvenile apparent survival<br />

obtained from <strong>the</strong> CJS model with banding data<br />

from juvenile spotted owls will probably not be<br />

useful from <strong>the</strong> pooled quadrat survey data<br />

because <strong>the</strong> emigration rate of this population<br />

segment will be quite high as <strong>the</strong>y disperse away<br />

from <strong>the</strong>ir natal territories.<br />

Personnel Personnel<br />

Personnel<br />

Quality work cannot be completed without<br />

capable people who desire to per<strong>for</strong>m well. <strong>Owl</strong><br />

surveys are difficult to conduct. To achieve<br />

accurate survey results requires a certain combination<br />

of physical and mental traits. The ideal<br />

candidate <strong>for</strong> spotted owl survey work must be<br />

physically capable of negotiating difficult terrain<br />

and doing so after dark. The mental demands<br />

include <strong>the</strong> intellectual capacity to understand<br />

<strong>the</strong> nuances of <strong>the</strong> work, <strong>the</strong> perseverance to<br />

succeed under adverse conditions, <strong>the</strong> ability to<br />

follow directions, and <strong>the</strong> discipline to be


patient. Rigorous training to certify people to<br />

conduct monitoring is a critical step to ensure<br />

that qualified people implement <strong>the</strong> procedures.<br />

Restrictions on duration of work day, night<br />

time work, and camping near survey sites can<br />

lead to inefficiency. Effective inventory and<br />

monitoring may often require personnel to<br />

survey between dusk and 2300 hrs and prior to<br />

sunrise <strong>the</strong> next morning if an owl is detected<br />

<strong>the</strong> previous night. Not permitting camping at<br />

sites or not allowing more than 8-hour work<br />

days is an ineffective survey strategy. Thus, cost<br />

and ef<strong>for</strong>t <strong>for</strong> determining occupancy or reproduction<br />

in a given territory may be doubled or<br />

tripled. Not allowing personnel to survey along<br />

marked ridge lines at night (i.e., off roads) may<br />

result in inadequate survey of an area. Competent,<br />

qualified, eager personnel can conduct such<br />

activities safely and with desired results as demonstrated<br />

by May et al. (in press).<br />

Training<br />

Training<br />

Training is <strong>the</strong> most important mechanism<br />

<strong>for</strong> ensuring quality data and standardization.<br />

The current certification program employed by<br />

<strong>the</strong> FS should continue, but in a more intense<br />

fashion. High-quality photographic media,<br />

including video tapes of proper procedures,<br />

should be incorporated. Use of a map, compass,<br />

and GPS system, and recording spatial in<strong>for</strong>mation<br />

with Universal Transverse Mercator (UTM)<br />

coordinates (Grubb and Eakle 1988) are critical.<br />

Additional training <strong>for</strong> recording in<strong>for</strong>mation on<br />

data <strong>for</strong>ms, maps, and in an electronic data base<br />

is required (see below). Certified owl biologists<br />

should be tested routinely on <strong>the</strong>ir ability to<br />

complete data <strong>for</strong>ms and plot locations correctly.<br />

A standardized procedure <strong>for</strong> storing all in<strong>for</strong>mation<br />

also needs to be developed and en<strong>for</strong>ced.<br />

All training must be rein<strong>for</strong>ced with adequate<br />

(4-day minimum) field exercises followed<br />

by periodic rein<strong>for</strong>cement of learned skills.<br />

Although initial skills may be provided with <strong>the</strong><br />

certification process, rein<strong>for</strong>cement through<br />

feedback on procedures and results is required.<br />

An electronic data entry program will help to<br />

standardize inputs and rein<strong>for</strong>ce proper documentation<br />

procedures. Such a routine will also<br />

indicate progress of <strong>the</strong> monitoring program and<br />

Volume I/Part III<br />

112<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

identify personnel or administrative units that<br />

need additional training. Fur<strong>the</strong>r, additional<br />

training should include periodic visits by program<br />

supervisors to review field procedures.<br />

Incorrect observations may not necessarily be<br />

detected on data <strong>for</strong>ms.<br />

We also suggest a greater emphasis on<br />

identification of spotted owl age classes (juvenile,<br />

subadult, and adult) as described by Forsman<br />

(1981) and Moen et al. (1991). This valuable<br />

in<strong>for</strong>mation may be obtained if observers take<br />

binoculars on surveys. In<strong>for</strong>mation on age<br />

structure may prove useful <strong>for</strong> identifying<br />

changes in demographic trends.<br />

All survey routes and results need to be<br />

summarized in a standardized manner on media<br />

that can be entered into a Geographic In<strong>for</strong>mation<br />

System (GIS). Thus, field personnel will<br />

require training on use of map, compass, and<br />

GPS. In addition, a standardized map system<br />

and symbol set is paramount. We recommend a<br />

7.5" USGS topographic map. This map type is<br />

readily available in paper and digital <strong>for</strong>m. We<br />

also insist that field personnel record UTM<br />

coordinates. This will allow rapid updating of<br />

digital maps of owl locations.<br />

Computerized Computerized Data Data Data Entry Entry and and<br />

and<br />

Summarization<br />

Summarization<br />

A major weakness of past inventory and<br />

monitoring programs has been <strong>the</strong> lack of<br />

accessibility to data; as a result, few summaries<br />

and analyses were prepared. This scarcity of data<br />

examination appears to be due to <strong>the</strong> lack of a<br />

central, accessible, computerized database where<br />

field <strong>for</strong>ms are regularly entered <strong>for</strong> computer<br />

analysis. Field workers submitting data <strong>for</strong>ms but<br />

not receiving feedback from <strong>the</strong>ir ef<strong>for</strong>ts nor a<br />

copy of <strong>the</strong> master database <strong>for</strong> <strong>the</strong>m to review<br />

leads to errors that are difficult to rectify retroactively.<br />

We suggest that field workers who collect<br />

<strong>the</strong> data should also be responsible <strong>for</strong> data entry<br />

into a standardized computer <strong>for</strong>m. The benefits<br />

would be twofold.<br />

First, a computerized data entry <strong>for</strong>m would<br />

guarantee that only admissible codes are used<br />

because invalid codes would not be accepted by<br />

<strong>the</strong> computer, and correct entries would be


needed in all data fields be<strong>for</strong>e <strong>the</strong> user could<br />

proceed. Quality control would be facilitated via<br />

an interactive computer data entry interface.<br />

With such a data entry program, data from<br />

different jurisdictions of all involved land management<br />

entities would be compatible.<br />

Second, once <strong>the</strong> data have been entered,<br />

summaries can be produced with standard<br />

summary programs. At a minimum, field workers<br />

should be able to produce summaries of data<br />

<strong>the</strong>y entered, and make comparisons with past<br />

years and maybe o<strong>the</strong>r geographic areas. The<br />

main reason <strong>for</strong> this instant feedback is to<br />

encourage field personnel to examine <strong>the</strong>ir own<br />

data plus get a temporal and spatial perspective<br />

of existing data. Field workers would have a<br />

much better picture of how <strong>the</strong>ir data fit into <strong>the</strong><br />

overall ef<strong>for</strong>t and would have access to data in<br />

<strong>the</strong> master database. Simple graphs and tabular<br />

summaries should be available via a menu<br />

system. This feedback would also promote<br />

greater cooperation in future surveys and makes<br />

<strong>the</strong> field worker feel a part of <strong>the</strong> complete<br />

process.<br />

Creation of a master database on an accessible<br />

computer network (such as <strong>the</strong> World Wide<br />

Web [WWW] on Internet) has ano<strong>the</strong>r, less<br />

apparent, benefit <strong>for</strong> <strong>the</strong> program. From this<br />

master database, region-wide summaries could<br />

be generated. Annual summaries would help<br />

detect trends in <strong>the</strong> data. Sophisticated statistical<br />

analyses could be programmed to implement<br />

tests <strong>for</strong> trends in <strong>the</strong> data. Safeguards would be<br />

necessary to limit access to <strong>the</strong> data, particularly<br />

sensitive site locations, to only authorized<br />

persons. Finally, scrutiny by outside reviewers<br />

would improve <strong>the</strong> integrity of <strong>the</strong> database.<br />

To implement <strong>the</strong> above scheme, two pieces<br />

of software need to be written. The first is <strong>the</strong><br />

data entry system, <strong>for</strong> which extensive error<br />

checking should be coded into <strong>the</strong> software.<br />

Data entered by a field worker would be appended<br />

to <strong>the</strong> master data file only after passing<br />

a stringent series of integrity checks. The second<br />

is a data summary program that would be menu<br />

driven and allow <strong>the</strong> user to summarize his/her<br />

own data plus o<strong>the</strong>r data of interest. We presume<br />

that modern PC computer software systems and<br />

access to a computer network should make this<br />

software development fairly easy. Once <strong>the</strong> field<br />

Volume I/Part III<br />

113<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

season is completed, each land management<br />

entity and <strong>the</strong>ir respective subunits should be<br />

able to obtain graphical summaries as well as<br />

statistical summaries in tabular <strong>for</strong>m.<br />

Costs<br />

Costs<br />

The cost <strong>for</strong> implementing <strong>the</strong> population<br />

monitoring scheme should include hiring a<br />

principal investigator to design this survey and<br />

coordinate sampling ef<strong>for</strong>ts. Field crew leaders<br />

will be necessary <strong>for</strong> supervising study logistics<br />

and field technicians will be required to conduct<br />

surveys. In addition, while models exist <strong>for</strong><br />

estimating total population size through time,<br />

models of multiple capture probabilities require<br />

some independent work. All field personnel<br />

hired to conduct <strong>the</strong> pilot study and subsequent<br />

monitoring program must be qualified and<br />

trained.<br />

Our initial estimate of <strong>the</strong> costs to fully<br />

implement <strong>the</strong> proposed monitoring scheme is<br />

approximately $1.2-1.5 million per year. Based<br />

on <strong>the</strong> delisting criteria, monitoring must<br />

continue <strong>for</strong> a minimum of 15 years.<br />

Costs <strong>for</strong> implementing macrohabitat<br />

monitoring are unknown. However, much of <strong>the</strong><br />

needed remote-sensing coverage exists or is being<br />

obtained as a tool <strong>for</strong> implementing ecosystem<br />

management. Thus, additional costs attributable<br />

to <strong>the</strong> spotted owl should be minimal. Costs of<br />

implementing microhabitat sampling are difficult<br />

to estimate without knowledge of <strong>the</strong><br />

number of plots to be sampled. We assume,<br />

however, that sampling an area pre- and posttreatment<br />

is already required as a standard part<br />

of activity implementation; consequently, <strong>the</strong><br />

cost attributable to owl monitoring would entail<br />

those associated with measuring additional<br />

variables specific to <strong>the</strong> owl. Thus, <strong>the</strong> total costs<br />

of habitat monitoring should be relatively<br />

minimal beyond that already required or in <strong>the</strong><br />

process of being developed independent of <strong>the</strong><br />

spotted owl.<br />

Potential Potential Experiments<br />

Experiments<br />

Many habitat variables important to <strong>Mexican</strong><br />

spotted owls cannot be monitored by remote<br />

sensing. Fur<strong>the</strong>r, it is important to ensure that


adequate habitat is provided <strong>for</strong> key prey as well.<br />

Thus, we propose some potential experiments to<br />

relate habitat conditions to owl population<br />

dynamics where key habitat characteristics would<br />

be measured on <strong>the</strong> ground. On-<strong>the</strong>-ground<br />

monitoring of relevant habitat characteristics<br />

would quantify <strong>the</strong>ir change at a local (i.e.,<br />

within quadrat) scale and relate <strong>the</strong>m to owl<br />

population dynamics.<br />

Population monitoring based on randomly<br />

selected quadrats provides <strong>the</strong> opportunity to<br />

conduct experiments to extend our knowledge of<br />

<strong>the</strong> impact of habitat manipulation on <strong>Mexican</strong><br />

spotted owl population dynamics. We propose<br />

<strong>the</strong>se experiments to produce credible, defensible,<br />

and reliable results (sensu Murphy and<br />

Noon 1991). Quadrats within <strong>the</strong> monitoring<br />

design may serve as experimental units <strong>for</strong><br />

examining <strong>the</strong> effects of future management<br />

such as fires, grazing, timber harvest, and recreation.<br />

Given that a treatment is identified prior to<br />

its occurrence, vegetation measurements can take<br />

place on <strong>the</strong> site of <strong>the</strong> expected treatment and a<br />

second, control quadrat that is selected based on<br />

its similarity to <strong>the</strong> expected treatment quadrat.<br />

This experimental design is not a true experiment,<br />

because <strong>the</strong> treatment is not randomly<br />

allocated to one of <strong>the</strong> pair of quadrats. However,<br />

this quasi-experiment is still more powerful<br />

in developing cause-and-effect relationships<br />

between habitat manipulations and owl population<br />

dynamics than <strong>the</strong> more common correlative<br />

designs used by past researchers (see III.D<br />

<strong>for</strong> fur<strong>the</strong>r details on experimental design).<br />

Fur<strong>the</strong>r, <strong>the</strong> capability to replicate <strong>the</strong> treatment<br />

exists because of <strong>the</strong> extensive number of quadrats<br />

that will be required <strong>for</strong> measuring changes<br />

in population size.<br />

Areas where planned treatments result in<br />

some <strong>for</strong>m of habitat alteration provide excellent<br />

opportunities <strong>for</strong> quasi-experiments. Vegetation<br />

measures should be taken immediately be<strong>for</strong>e<br />

and after <strong>the</strong> habitat-modifying event, and<br />

<strong>the</strong>reafter at 5-year intervals. Vegetation measurements<br />

that seem especially important to<br />

examine are tree size-class distribution, log sizeclass<br />

distribution, canopy cover, and shrub cover.<br />

Results from <strong>the</strong>se experiments, coupled with<br />

results of population monitoring, will provide<br />

Volume I/Part III<br />

114<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>the</strong> basis <strong>for</strong> a predictive model of spotted owl<br />

habitat quality (assuming that owl density<br />

reflects habitat quality). Data on apparent owl<br />

survival and reproduction will also be available,<br />

which may relate to habitat quality more directly<br />

than owl density.<br />

Alternative Alternative Designs Designs Designs <strong>for</strong> <strong>for</strong> Population<br />

Population<br />

Monitoring<br />

Monitoring<br />

Drawing Drawing New New Sample Sample of of Quadrats Quadrats Quadrats Each Each Each Year<br />

Year<br />

Instead of drawing an initial sample of<br />

quadrats from <strong>the</strong> sampling frame and monitoring<br />

<strong>the</strong>se same quadrats through time, an alternative<br />

approach would be to draw a completely<br />

new random sample of quadrats each year. For<br />

repeated sampling of a set of quadrats to be<br />

legitimate, normal activities that occur in spotted<br />

owl habitat should continue during <strong>the</strong><br />

monitoring program, provided <strong>the</strong>se activities<br />

meet <strong>the</strong> requirements of section 7(a)(2) of <strong>the</strong><br />

Act by not likely jeopardizing <strong>the</strong> continued<br />

existence of <strong>the</strong> <strong>Mexican</strong> spotted owl. The main<br />

advantage of a new sample each year is that it<br />

guards against <strong>the</strong> potential <strong>for</strong> land managers to<br />

manage areas within <strong>the</strong> quadrats differently<br />

than <strong>the</strong> remainder of <strong>the</strong> landscape. The price<br />

of this protection is relatively great as illustrated<br />

by <strong>the</strong>se four points: (1) <strong>the</strong> logistics of conducting<br />

<strong>the</strong> surveys each year would increase because<br />

of <strong>the</strong> new quadrats; (2) age-specific apparent<br />

survival rates and recruitment to <strong>the</strong> territorial<br />

population could not be estimated with <strong>the</strong> CJS<br />

analysis because birds would not be marked on<br />

<strong>the</strong> same area each year; (3) quasi-experiments to<br />

detect <strong>the</strong> relationship between habitat manipulations<br />

and owl population dynamics would not<br />

be possible; and (4) higher sampling intensities<br />

would be required because this design is less<br />

efficient <strong>for</strong> estimating change. Our proposed<br />

design is intermediate between sampling <strong>the</strong><br />

same set of quadrats each year and a completely<br />

new sample each year. We obtain <strong>the</strong> benefits<br />

from both alternatives in that <strong>the</strong> correlation of<br />

measurements <strong>for</strong> a specific quadrat across years<br />

is used to lower <strong>the</strong> overall variance of our<br />

population estimate, making <strong>the</strong> design more<br />

efficient than complete replacement each year,<br />

yet we are guarding against <strong>the</strong> potential <strong>for</strong>


sampled quadrats to be managed differently than<br />

o<strong>the</strong>r areas.<br />

Conducting Conducting Surveys Surveys Less Less Often Often Than Than Yearly<br />

Yearly<br />

Instead of surveying quadrats each year,<br />

ef<strong>for</strong>t and cost could be saved by conducting <strong>the</strong><br />

surveys at longer intervals, such as every 5 years.<br />

An advantage of this approach is that costs will<br />

be lowered, and possibly more precise estimates<br />

of population size could be obtained by pooling<br />

money to conduct a few very good surveys<br />

instead of more frequent surveys with lower<br />

ef<strong>for</strong>t per survey. The main disadvantage of this<br />

approach is that age-specific apparent survival<br />

rates and recruitment to <strong>the</strong> territorial population<br />

could not be estimated with <strong>the</strong> CJS analysis<br />

because birds would not be marked frequently<br />

enough to obtain <strong>the</strong>se estimates. For example,<br />

given an estimate of 0.89 <strong>for</strong> adult survival, only<br />

56% of <strong>the</strong> initial population would still be alive<br />

after 5 years, resulting in small sample sizes of<br />

recaptured birds, and hence poorer precision of<br />

<strong>the</strong> survival estimates. Fur<strong>the</strong>r, reproductive and<br />

annual survival rates and <strong>the</strong>ir variation across<br />

years are needed to realistically evaluate population<br />

viability. Finally, our ability to detect<br />

relationships between habitat manipulations and<br />

population dynamics would be greatly decreased<br />

because this approach is more sensitive to variability<br />

introduced by <strong>the</strong> years chosen <strong>for</strong> sampling.<br />

Volume I/Part III<br />

115<br />

Adaptive Adaptive Sampling<br />

Sampling<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Thompson (1992) has developed an adaptive<br />

sampling scheme to improve <strong>the</strong> efficiency of<br />

sampling clustered populations, such as is<br />

probably <strong>the</strong> case <strong>for</strong> <strong>Mexican</strong> spotted owls.<br />

Thompson’s scheme is <strong>the</strong>oretically appealing<br />

because more ef<strong>for</strong>t is applied to areas where<br />

owls are located. Under this approach, quadrats<br />

adjoining a quadrat that contains some threshold<br />

number of spotted owls would also be sampled.<br />

Un<strong>for</strong>tunately, we cannot envision how to<br />

handle <strong>the</strong> logistics of adding some unknown<br />

number of quadrats to <strong>the</strong> sample when survey<br />

crews must be hired, trained, and outfitted with<br />

equipment and vehicles prior to sampling. We<br />

suspect <strong>the</strong> logistical overhead of this approach<br />

may make it impractical <strong>for</strong> monitoring owls on<br />

quadrats. However, as <strong>the</strong> <strong>the</strong>ory and application<br />

of <strong>the</strong> adaptive sampling scheme is developed<br />

fur<strong>the</strong>r, an innovative application of <strong>the</strong> technique<br />

may be possible with our proposed quadrat<br />

monitoring scheme.<br />

CONCLUSION<br />

CONCLUSION<br />

The technology and expertise are available to<br />

monitor trends in <strong>Mexican</strong> spotted owl habitat<br />

and population size. Clearly, <strong>the</strong> objectives and<br />

design of <strong>the</strong> monitoring program must be<br />

defined explicitly and <strong>the</strong>y must be attainable.<br />

To implement <strong>the</strong> process, knowledgeable,<br />

dedicated people must be assigned <strong>the</strong> task.<br />

Adequate training and constant feedback mechanisms<br />

are critical aspects to a successful monitoring<br />

program as tenable conclusions can be based<br />

only on reliable data.


The primary objectives of our proposed<br />

research program are to (1) enhance understanding<br />

of <strong>Mexican</strong> spotted owl biology and (2)<br />

assess how land management practices affect <strong>the</strong><br />

owl population’s viability. These types of in<strong>for</strong>mation<br />

are necessary to complement recovery<br />

ef<strong>for</strong>ts outlined in this plan. The research program<br />

described here is different from <strong>the</strong> monitoring<br />

program outlined in III.C. Whereas both<br />

programs are necessary, specific research needs<br />

may or may not be related to monitoring. In<br />

developing this chapter, we realized that readers<br />

of this plan have a variety of backgrounds. Thus,<br />

to establish a common framework <strong>for</strong> <strong>the</strong> discussion<br />

of a research program <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, we first outline <strong>the</strong> role of <strong>the</strong><br />

scientific process in research and some important<br />

aspects of study design. We <strong>the</strong>n discuss some<br />

limitations with previous research, and suggest<br />

future research questions and processes that<br />

should be examined.<br />

ROLE ROLE OF OF THE THE SCIENTIFIC<br />

SCIENTIFIC<br />

PROCESS<br />

PROCESS<br />

Research and <strong>the</strong> reliability of knowledge<br />

gained from research depend on appropriate<br />

application of <strong>the</strong> scientific method. Reliable<br />

knowledge can be defined as “<strong>the</strong> set of ideas<br />

that agree or are consistent with <strong>the</strong> facts of<br />

nature,” whereas “unreliable knowledge is <strong>the</strong> set<br />

of false ideas mistaken <strong>for</strong> knowledge”<br />

(Romesburg 1981). Three primary scientific<br />

methods have been used in scientific research<br />

(Romesburg 1981): (1) induction that involves<br />

<strong>the</strong> use of repeated observations to discover laws<br />

of association; (2) retroduction where a “bestguess”<br />

hypo<strong>the</strong>sis is developed to explain a law of<br />

association or some set of observations; and (3)<br />

hypo<strong>the</strong>tico-deductive (HD) where a priori hypo<strong>the</strong>ses<br />

are developed and tested, and a decision<br />

made about whe<strong>the</strong>r to reject <strong>the</strong> hypo<strong>the</strong>ses. It<br />

is generally accepted in science that application<br />

of <strong>the</strong> HD method provides <strong>the</strong> best avenue <strong>for</strong><br />

gaining reliable knowledge (Platt 1964, Popper<br />

1965, Romesburg 1981, 1991). Steps used in<br />

Volume I/Part III<br />

D. D. D. ACTIVITY-SPECIFIC<br />

ACTIVITY-SPECIFIC<br />

RESEARCH<br />

RESEARCH<br />

116<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>the</strong> HD method can be reiterated as follows<br />

from Nichols (1991): “(1) suggest a hypo<strong>the</strong>sis<br />

to explain some phenomenon of interest, (2)<br />

deduce a testable prediction from that hypo<strong>the</strong>sis,<br />

(3) devise and carry out a suitable test, and<br />

(4) use observations from <strong>the</strong> test to decide<br />

whe<strong>the</strong>r <strong>the</strong> prediction is met.” When observations<br />

and predictions match, <strong>the</strong> hypo<strong>the</strong>sis is<br />

corroborated; when <strong>the</strong>y do not match, <strong>the</strong><br />

hypo<strong>the</strong>sis has been falsified and can be discarded.<br />

Rejection of hypo<strong>the</strong>ses is key to <strong>the</strong> HD<br />

method. Corroboration of hypo<strong>the</strong>ses can result<br />

from poor experimental designs (e.g., low<br />

power). There<strong>for</strong>e, knowledge in <strong>the</strong> HD<br />

method is gained more through falsification of<br />

hypo<strong>the</strong>ses than through corroboration.<br />

Most research related to natural resource<br />

management and conservation has relied primarily<br />

on induction and retroduction (Romesburg<br />

1981, 1991). Induction can provide us with<br />

reliable knowledge about associations such as <strong>the</strong><br />

association of <strong>Mexican</strong> spotted owls with <strong>for</strong>ests<br />

having certain structural characteristics. However,<br />

this method does not provide <strong>the</strong> mechanism<br />

<strong>for</strong> understanding <strong>the</strong> processes that<br />

underlie this association nor does it provide<br />

reliable knowledge about cause and effect.<br />

Whereas we can describe <strong>the</strong> structure of <strong>for</strong>ests<br />

used by spotted owls, we cannot ascertain which<br />

structural characteristics are “important,” or why,<br />

without application of <strong>the</strong> HD method. In<br />

short, we can describe patterns through induction<br />

but need <strong>the</strong> HD method to understand<br />

why those patterns occur and which components<br />

of those patterns are “important.” In terms of<br />

management, understanding why a pattern has<br />

occurred and what caused it are important <strong>for</strong><br />

predicting effects when observed patterns are<br />

changed.<br />

Romesburg (1981) argued that retroduction<br />

does not provide reliable knowledge because of<br />

<strong>the</strong> inability of this method to falsify hypo<strong>the</strong>ses<br />

and <strong>the</strong> large number of alternative hypo<strong>the</strong>ses<br />

that could equally explain <strong>the</strong> same conclusions.<br />

However, both induction and retroduction are<br />

useful <strong>for</strong> describing relationships and develop-


ing hypo<strong>the</strong>ses to be fur<strong>the</strong>r tested using <strong>the</strong> HD<br />

method. Unreliability of retroduction can be<br />

exacerbated when untested hypo<strong>the</strong>ses are<br />

integrated into our knowledge base as dogma in<br />

<strong>the</strong> <strong>for</strong>m of scientific “rules.”<br />

While induction and <strong>the</strong> HD method<br />

provide a general framework <strong>for</strong> gaining reliable<br />

knowledge, design of appropriate studies is<br />

crucial to <strong>the</strong> application of this method in<br />

specific situations. This applies to both describing<br />

and understanding patterns in nature. Any<br />

management plan, including <strong>the</strong> <strong>Mexican</strong><br />

<strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>, is a complex hypo<strong>the</strong>sis<br />

whose rejection or corroboration is<br />

determined by <strong>the</strong> success or failure of <strong>the</strong> plan<br />

over <strong>the</strong> long term.<br />

IMPORTANT IMPORTANT ASPECTS ASPECTS OF<br />

OF<br />

EXPERIMENTAL EXPERIMENTAL EXPERIMENTAL DESIGN<br />

DESIGN<br />

The ability to confidently infer results from a<br />

sample to a population of interest and <strong>the</strong><br />

strength of that inference are entirely dependent<br />

on study design. Reliability of knowledge and<br />

<strong>the</strong> ability to make correct inferences are directly<br />

proportional; <strong>the</strong> stronger <strong>the</strong> inference one can<br />

make, <strong>the</strong> more reliable <strong>the</strong> knowledge stemming<br />

from that inference. The strongest inference<br />

in understanding patterns is achieved<br />

through controlled experiments, with <strong>the</strong><br />

strength of inference diminishing <strong>the</strong> fur<strong>the</strong>r a<br />

given study design departs from <strong>the</strong> experimental<br />

(HD) approach. However, inferences can be<br />

weakened even in experimental studies if <strong>the</strong><br />

design is not valid. With <strong>Mexican</strong> spotted owls,<br />

we would like to extend inferences to a larger<br />

population than <strong>the</strong> one from which we<br />

sampled. This larger population may be across<br />

<strong>the</strong> range of <strong>the</strong> owl, within a certain recovery<br />

unit, or on a Ranger District within a National<br />

Forest. The ability to extend conclusions from a<br />

study to a larger area or a longer time period<br />

depends directly on how <strong>the</strong> study was designed<br />

and implemented.<br />

For our purposes, study designs can be<br />

characterized as ei<strong>the</strong>r descriptive or experimental<br />

(following Eberhardt and Thomas 1991).<br />

Descriptive studies employ survey sampling,<br />

whereas experimental studies use treatment and<br />

Volume I/Part III<br />

117<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

control groups. Necessary components of <strong>the</strong><br />

design in both cases include: (1) randomization<br />

where samples are randomly selected in a descriptive<br />

study, or treatments and controls are<br />

randomly assigned in an experiment; and (2)<br />

replication of experimental units through both<br />

space and time. Randomization removes subjective<br />

biases that may be found in descriptive<br />

studies and guards against systematic differences<br />

o<strong>the</strong>r than treatment effects in experiments.<br />

Randomization also allows <strong>for</strong> stronger inference<br />

to a larger population and is <strong>the</strong> <strong>the</strong>oretical basis<br />

<strong>for</strong> employing statistical tests. Replication allows<br />

<strong>for</strong> estimation of experimental error, a prerequisite<br />

<strong>for</strong> employing statistical tests. If ei<strong>the</strong>r<br />

randomization or replication are omitted from<br />

an experimental design, inferences will be greatly<br />

weakened. True replication should not be confused<br />

with “pseudoreplication” where<br />

subsampling of experimental units is confused<br />

with replication of experimental units (Hurlbert<br />

1984). For example, a habitat study that measures<br />

100 vegetation plots within each of four<br />

owl home ranges represents a sample of four, not<br />

400. Frequently, researchers are guilty of<br />

pseudoreplication by reporting a sample size of<br />

400. Inferences from such a study apply only to<br />

<strong>the</strong> 4 owls studied and not to a larger population.<br />

Thus, adequate replication must occur at<br />

<strong>the</strong> level of <strong>the</strong> experimental unit (in this case,<br />

number of home ranges) to apply to a larger<br />

population.<br />

A major difficulty in doing field experiments<br />

is that <strong>the</strong>y are per<strong>for</strong>med in an uncontrolled,<br />

“noisy” environment (Eberhardt and Thomas<br />

1991). There<strong>for</strong>e, pre- and post-treatment<br />

measurement periods in both control and<br />

treatment groups are needed to reduce <strong>the</strong> effects<br />

of external variation and to ensure that a treatment<br />

effect can be adequately measured. Additional<br />

important design features necessary in<br />

field experiments include <strong>the</strong> choice of experimental<br />

units (e.g., owls, owl sites), local control<br />

(amount of balancing and blocking of experimental<br />

units), and <strong>the</strong> choice of <strong>the</strong> design (e.g.,<br />

complete block, incomplete block, factorial).<br />

An important consideration when designing<br />

and implementing studies that involve testing<br />

statistical hypo<strong>the</strong>ses is <strong>the</strong> power of <strong>the</strong> statistical<br />

test used (<strong>the</strong> probability of rejecting <strong>the</strong> null


hypo<strong>the</strong>sis when it is false). Failure to reject a<br />

null hypo<strong>the</strong>sis is due to ei<strong>the</strong>r (1) <strong>the</strong> null<br />

hypo<strong>the</strong>sis was indeed “true” or (2) <strong>the</strong>re was<br />

insufficient power to reject it. Thus, power<br />

should be as high as possible (>90%) to corroborate<br />

that an unfalsified null hypo<strong>the</strong>sis was<br />

actually not false. Power is dependent on a<br />

combination of <strong>the</strong> severity of <strong>the</strong> treatment<br />

applied, sample size, and experimental error. If a<br />

treatment is subtle, <strong>the</strong>n a larger sample will be<br />

necessary to achieve <strong>the</strong> same power as if a severe<br />

treatment was used. This is important because<br />

biological questions often involve chronic<br />

(subtle) effects ra<strong>the</strong>r than acute (severe) effects.<br />

For example, <strong>the</strong> effects of a given land management<br />

practice may have a slight effect on adult<br />

survival rates which may in turn have strong<br />

effects on population viability. If an experiment<br />

testing such an effect has low power, <strong>the</strong>n it may<br />

be tempting to state that <strong>the</strong> practice does not<br />

significantly affect survival rates when in fact it<br />

does. Repercussions from an experiment lacking<br />

sufficient power would <strong>the</strong>n be misleading and<br />

result in false confidence in <strong>the</strong> health of <strong>the</strong><br />

population.<br />

LIMITATIONS LIMITATIONS IN IN PAST<br />

PAST<br />

RESEARCH RESEARCH ON ON THE THE MEXICAN MEXICAN<br />

MEXICAN<br />

SPOTTED SPOTTED OWL<br />

OWL<br />

Previous research on <strong>Mexican</strong> spotted owls<br />

has been largely descriptive and has relied on<br />

induction and retroduction; our current knowledge<br />

concerning underlying ecological processes<br />

is, <strong>the</strong>re<strong>for</strong>e, limited. However, previous research<br />

on <strong>Mexican</strong> spotted owls has provided a good<br />

foundation to describe <strong>the</strong> natural history of <strong>the</strong><br />

species and to generate hypo<strong>the</strong>ses <strong>for</strong> experimental<br />

tests with <strong>the</strong> HD method. Additional<br />

limitations on conclusions from previous research<br />

result from (1) lack of randomization in<br />

selecting experimental units and study areas, (2)<br />

lack of true replication (including small sample<br />

sizes), and (3) lack of experiments. The following<br />

discussion is not meant as criticism of<br />

specific research studies or scientists. Many of<br />

<strong>the</strong> previous studies have been hampered by<br />

inadequate funding and logistical constraints<br />

beyond <strong>the</strong> investigators’ control. These factors<br />

Volume I/Part III<br />

118<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

are not unique to research on <strong>the</strong> <strong>Mexican</strong><br />

spotted owl; <strong>the</strong>y are common to research on<br />

numerous species, including both <strong>the</strong> nor<strong>the</strong>rn<br />

and Cali<strong>for</strong>nia spotted owls.<br />

Lack of randomization and replication has<br />

hampered <strong>the</strong> ability to infer from particular<br />

samples to <strong>the</strong> general. In a number of studies,<br />

pseudoreplication has also been confused with<br />

true replication, weakening inferences even<br />

fur<strong>the</strong>r. For example, most of <strong>the</strong> habitat studies<br />

using radiotelemetry have suffered from<br />

pseudoreplication. These studies typically<br />

sampled few (4-10) birds, but sampled habitat<br />

characteristics within <strong>the</strong>se few birds’ home<br />

ranges extensively. In testing hypo<strong>the</strong>ses, <strong>the</strong><br />

number of subsamples were used, ra<strong>the</strong>r than <strong>the</strong><br />

number of owls, to estimate error terms used in<br />

statistical tests. Such pseudoreplication lends an<br />

incorrect perception of adequate power to<br />

statistical tests which may lead to incorrect<br />

conclusions. However, repetition of home-range<br />

studies over additional areas has streng<strong>the</strong>ned<br />

inferences concerning certain habitat associations.<br />

Controlled experiments have not been used<br />

in research on <strong>Mexican</strong> spotted owls. Lack of<br />

experiments is probably related to <strong>the</strong> need to<br />

quickly identify basic aspects of spotted owl<br />

natural history and apply this in<strong>for</strong>mation to<br />

management situations. However, experiments<br />

are critical <strong>for</strong> defining <strong>the</strong> impacts of current<br />

and proposed management activities on <strong>Mexican</strong><br />

spotted owls.<br />

RESEARCH RESEARCH NEEDS<br />

NEEDS<br />

Several management issues and questions<br />

must be resolved to better understand and<br />

implement recovery measures <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Communication and collaboration<br />

between people with strong research skills and<br />

people with strong management skills will be a<br />

key component in this process. Managers need<br />

to better understand <strong>the</strong> methods, problems and<br />

uncertainties involved with research. Researchers,<br />

on <strong>the</strong> o<strong>the</strong>r hand, must rely on managers to<br />

identify appropriate questions, political and legal<br />

constraints, and to develop appropriate implementation<br />

of knowledge derived from research


esults. Too often researchers design and implement<br />

studies that do not adequately address<br />

management problems. People having both<br />

research and management skills will hopefully<br />

bridge <strong>the</strong> gap between management and research<br />

disciplines. Both time and money are<br />

short. Clearly, all research questions cannot be<br />

answered within a short time frame. There<strong>for</strong>e,<br />

we advocate that a series of crucial experiments<br />

be implemented that address questions most<br />

relevant to <strong>the</strong> needs of management agencies.<br />

The following example of such an experiment<br />

addresses <strong>the</strong> question, “what structural features<br />

in <strong>for</strong>est habitat are needed to maintain high<br />

fitness in <strong>Mexican</strong> spotted owls,” where fitness is<br />

some function of survival and reproduction:<br />

Volume I/Part III<br />

Determine appropriate statistical hypo<strong>the</strong>ses<br />

(predictions) and response<br />

variables to be tested. Testing fitness<br />

directly through survival and reproductive<br />

rates may not be feasible because of<br />

<strong>the</strong> prohibitively large samples needed to<br />

detect chronic effects and ethical problems<br />

in purposely affecting survival of<br />

<strong>the</strong> owls. However, appropriate hypo<strong>the</strong>ses<br />

from <strong>the</strong> initial question is that<br />

decline in <strong>for</strong>aging use and prey availability<br />

in altered habitats would directly<br />

affect fitness.<br />

Determine <strong>the</strong> extent and magnitude of<br />

treatments to apply to <strong>for</strong>ested habitat.<br />

For example, testable research hypo<strong>the</strong>ses<br />

could be that <strong>the</strong> extent of large trees in<br />

sites affects <strong>for</strong>aging use by owls and prey<br />

abundance. Treatments may be nested so<br />

that <strong>the</strong> same experimental units can be<br />

used in repeated experiments, assuming<br />

that treatments can be decided upon<br />

be<strong>for</strong>ehand and applied consecutively. In<br />

addition, treatments need not be “negative”<br />

by removing habitat components<br />

but can be “positive” by treating previously<br />

impacted habitats. Thus, careful<br />

planning is needed at this stage.<br />

Randomly select n spotted owl sites so<br />

that sufficient power can be achieved to<br />

detect differences in <strong>for</strong>aging by owls<br />

119<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

between treatments. Attach radiotransmitters<br />

to owls within sites.<br />

Collect pre-treatment data and define<br />

high-use areas by owls within all sites.<br />

Randomly assign treatment and control<br />

classifications to <strong>the</strong> n owl sites.<br />

Apply <strong>the</strong> treatment to high-use <strong>for</strong>aging<br />

areas within treatment sites only.<br />

Collect post-treatment data.<br />

Test <strong>for</strong> differences between treatment<br />

and control groups.<br />

Continue <strong>the</strong> same procedure with<br />

additional treatments.<br />

While not ideal, such an experiment illustrates<br />

<strong>the</strong> principles of scientific experimental<br />

design necessary to achieve reliable knowledge<br />

concerning spotted owl habitat use and, indirectly,<br />

fitness, as a function of <strong>for</strong>est structure.<br />

Such crucial experiments are difficult to design,<br />

require commitments of funding, and scientific<br />

imagination because of ethical constraints and<br />

<strong>the</strong> limitations on allowable habitat alterations<br />

proposed in this plan. However, <strong>the</strong>se types of<br />

experiments also more rapidly answer pressing<br />

management questions.<br />

We recommend research on <strong>the</strong> following<br />

questions about <strong>Mexican</strong> spotted owls that still<br />

need answers. Clearly, a large number of research<br />

questions could be developed that address all<br />

aspects of <strong>Mexican</strong> spotted owl biology <strong>for</strong><br />

which knowledge is lacking. However, we pose<br />

what we believe are <strong>the</strong> most crucial questions<br />

that need to be addressed in terms of immediate<br />

management problems and <strong>the</strong> recovery of <strong>the</strong><br />

owl. Studies designed to answer <strong>the</strong>se questions<br />

will be descriptive, experimental, or a combination<br />

of both.<br />

Dispersal<br />

Dispersal<br />

Dispersal is a key process in metapopulation<br />

<strong>the</strong>ory and to maintain genetic diversity between


isolated subpopulations (Keitt et al. 1995). Key<br />

questions include:<br />

Genetics<br />

Genetics<br />

Volume I/Part III<br />

Are subpopulations within and between<br />

<strong>Recovery</strong> Units connected?<br />

What habitats and large-scale habitat<br />

configurations do dispersing juveniles<br />

require to maintain adequate survival<br />

rates during dispersal?<br />

<strong>Mexican</strong> spotted owl populations are naturally<br />

fragmented across <strong>the</strong>ir range. Genetics can<br />

provide insight into historical connections<br />

between subpopulations. There<strong>for</strong>e, questions<br />

on genetics also relate to dispersal. Key questions<br />

include:<br />

Habitat Habitat<br />

Habitat<br />

Are subpopulations within and between<br />

<strong>Recovery</strong> Units genetically isolated and<br />

to what degree?<br />

What is <strong>the</strong> extent of genetic interchange<br />

across <strong>the</strong> entire range of <strong>the</strong> owl?<br />

<strong>Mexican</strong> spotted owls use a variety of<br />

habitats ranging from canyons to <strong>for</strong>ested areas<br />

(Ganey and Dick 1995). Key questions include:<br />

To what extent is habitat use determined<br />

by various factors, such as prey availability,<br />

temperature regulation, and/or<br />

avoidance of predators?<br />

What habitat components confer high<br />

fitness?<br />

How do land management activities,<br />

specifically grazing, timber harvest, fire,<br />

and recreation use, proximately affect<br />

habitat use and ultimately affect fitness?<br />

Population Population Biology<br />

Biology<br />

Currently, little is known about <strong>Mexican</strong><br />

spotted owl populations. Key questions can be<br />

addressed with our proposed monitoring plan:<br />

Is <strong>the</strong> <strong>Mexican</strong> spotted owl population<br />

stable, increasing, or declining?<br />

120<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Are some subpopulations increasing<br />

while o<strong>the</strong>rs are decreasing within <strong>the</strong><br />

range of <strong>the</strong> owl?<br />

Threats Threats to to <strong>Recovery</strong><br />

<strong>Recovery</strong><br />

Perceived threats need to be examined in<br />

relation to current management strategies to<br />

examine whe<strong>the</strong>r <strong>the</strong>se strategies are appropriate<br />

and to develop appropriate management strategies.<br />

Key questions include:<br />

What management strategies can be<br />

employed to reduce to possibility of<br />

catastrophic loss of owl habitat by fire<br />

while maintaining important habitat<br />

components?<br />

To what extent does disturbance from<br />

recreation, vehicles, etc. affect use of sites<br />

by spotted owls?<br />

How does grazing affect prey abundance<br />

in habitats used by spotted owls <strong>for</strong><br />

<strong>for</strong>aging?<br />

O<strong>the</strong>r O<strong>the</strong>r Ecosystem Ecosystem Components<br />

Components<br />

Implementation of <strong>the</strong> recovery measures <strong>for</strong><br />

<strong>the</strong> <strong>Mexican</strong> spotted owls will directly and<br />

indirectly affect numerous ecosystem attributes.<br />

Research is needed to determine <strong>the</strong> extent of<br />

<strong>the</strong>se effects on biotic and abiotic components,<br />

and ecosystem processes and function. Key<br />

questions are:<br />

What are <strong>the</strong> effects of this recovery plan<br />

on o<strong>the</strong>r vertebrates?<br />

What are <strong>the</strong> effects of implementing <strong>the</strong><br />

<strong>Plan</strong> on nonvertebrates?<br />

What are <strong>the</strong> effects of implementing <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> on plant community<br />

structure and composition?<br />

What are <strong>the</strong> effects of implementing <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> on abiotic ecosystem<br />

processes (e.g., hydrological systems)?<br />

What are <strong>the</strong> effects of implementing <strong>the</strong><br />

plan on ecosystem structure and function?<br />

How might <strong>the</strong> recovery plan be adjusted<br />

to mitigate potentially deleterious effects<br />

on o<strong>the</strong>r ecosystem attributes?


The ultimate goal of this <strong>Recovery</strong> <strong>Plan</strong> is<br />

to “recover” <strong>the</strong> <strong>Mexican</strong> spotted owl from<br />

threatened status. This action is referred to as<br />

“delisting” and is governed by section 4 of <strong>the</strong><br />

Act. Delisting <strong>the</strong> <strong>Mexican</strong> spotted owl will<br />

require reexamination of <strong>the</strong> same five factors<br />

considered during every listing process. In<br />

addition, five specific criteria have been developed<br />

to aid <strong>the</strong> delisting determination. Three of<br />

<strong>the</strong>se criteria pertain to <strong>the</strong> entire range of <strong>the</strong><br />

owl and two refer to a recovery unit level. The<br />

rangewide delisting criteria are:<br />

1. The populations in <strong>the</strong> Upper Gila<br />

Mountains, Basin and Range-East, and<br />

Basin and Range - West RUs must be<br />

shown to be stable or increasing after 10<br />

years of monitoring, using a study design<br />

with a power of 90% to detect a 20%<br />

decline with a Type I error rate of 0.05.<br />

2. Scientifically-valid habitat monitoring<br />

protocols are designed and implemented<br />

to assess (a) gross changes in habitat<br />

quantity across <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, and (b) whe<strong>the</strong>r microhabitat<br />

modifications and trajectories within<br />

treated stands meet <strong>the</strong> intent of <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong>.<br />

3. A long-term, U.S.-rangewide management<br />

plan is in place to ensure appropriate<br />

management of <strong>the</strong> subspecies and<br />

adequate regulation of human activity<br />

over time.<br />

Once <strong>the</strong> above three criteria are met,<br />

delisting may occur in any RU that meets <strong>the</strong><br />

final two criteria:<br />

4. Threats to <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

within <strong>the</strong> RU are sufficiently moderated<br />

and/or regulated.<br />

5. Habitat of a quality to sustain persistent<br />

<strong>Mexican</strong> spotted owl populations is<br />

stable or increasing within <strong>the</strong> RU.<br />

Volume I/Part III<br />

E. E. SUMMARY SUMMARY OF OF RECOVERY<br />

RECOVERY<br />

121<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>Recovery</strong> of <strong>the</strong> <strong>Mexican</strong> spotted owl hinges<br />

on successful implementation of three interrelated<br />

programs: population and habitat monitoring,<br />

management guidelines, and research.<br />

These aspects are not intended to stand alone;<br />

thus, all programs must be implemented simultaneously.<br />

For example, monitoring provides a<br />

measure of <strong>the</strong> effectiveness of <strong>the</strong> management<br />

guidelines. Without such monitoring, we will<br />

have no basis <strong>for</strong> determining whe<strong>the</strong>r management<br />

guidelines lead to <strong>the</strong> desired outcomes,<br />

and thus whe<strong>the</strong>r <strong>the</strong> bird should be delisted.<br />

Research is needed to answer key questions<br />

relevant to <strong>the</strong> <strong>Mexican</strong> spotted owl, particularly<br />

how implementation of management recommendations<br />

will affect <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

and its habitat. The knowledge derived from this<br />

research will provide a scientific basis <strong>for</strong> revising<br />

short-term guidelines and developing a longterm<br />

management plan.<br />

We have proposed a quadrat sampling<br />

scheme and provide detailed considerations <strong>for</strong><br />

determining spotted owl population trends<br />

within <strong>the</strong> Upper Gila Mountains, Basin and<br />

Range - East, and Basin and Range - West RUs.<br />

Population monitoring is not required <strong>for</strong> o<strong>the</strong>r<br />

recovery units because of sampling constraints<br />

posed by smaller population sizes. The suggested<br />

scheme provides a statistically valid means <strong>for</strong><br />

assessing population change Initial cost estimates<br />

<strong>for</strong> <strong>the</strong> owl monitoring scheme will range<br />

from $1.2 to $1.5 million per year.<br />

Habitat monitoring is needed to estimate<br />

trends in <strong>the</strong> quantity and quality of <strong>the</strong> owl’s<br />

habitat through time. Rangewide monitoring of<br />

<strong>the</strong> owl’s habitat should be conducted in conjunction<br />

with population monitoring. Because<br />

of <strong>the</strong> areal extent over which monitoring will be<br />

required, we propose <strong>the</strong> use of satellite imagery<br />

<strong>for</strong> tracking gross losses in habitat. We also<br />

propose that field sampling be conducted in<br />

conjunction with planned management treatments.<br />

Treatments include <strong>the</strong> use of prescribed<br />

fire, thinning, and silviculture. Monitoring<br />

should be done prior to and immediately following<br />

<strong>the</strong> treatment, and <strong>the</strong>n at five-year intervals.<br />

The objective of this sampling is to determine


changes to microhabitat features and also to<br />

verify that vegetation was placed or continues on<br />

a trajectory to become replacement habitat.<br />

Threats to be moderated include those that<br />

need site-specific treatment to alleviate <strong>the</strong>m,<br />

such as <strong>the</strong> reduction of <strong>the</strong> risks of catastrophic<br />

fire. For threats to be considered moderated,<br />

reasonable progress must have been made to<br />

remove identified threats and <strong>the</strong>re must be<br />

adequate assurance that management programs<br />

will continue. Long-term management plans are<br />

needed to guide management after <strong>the</strong> bird is<br />

delisted. Threats to be regulated include those<br />

resulting from agency management programs or<br />

o<strong>the</strong>r anthropogenic activities that are ei<strong>the</strong>r<br />

ongoing or reasonably certain to occur. These<br />

types of threats include wildfire hazard, timber<br />

harvest, urban or rural land development,<br />

grazing, and recreation.<br />

A primary focus of this <strong>Recovery</strong> <strong>Plan</strong> is to<br />

provide recommendations that will moderate or<br />

regulate threats over <strong>the</strong> short term (10-15<br />

years). Conceptually, this requires <strong>the</strong> presence<br />

of a mosaic of successional stages throughout a<br />

landscape comprised of <strong>the</strong> different habitats<br />

used by <strong>Mexican</strong> spotted owls. The arrangement<br />

and diversity of <strong>the</strong>se habitats must promote <strong>the</strong><br />

owl’s persistence. The short-term strategy is<br />

aimed at protecting existing owl habitat and<br />

initiating a process to develop replacement<br />

habitat. Although <strong>the</strong> approach is not completely<br />

synonymous with ecosystem management,<br />

implementation of <strong>the</strong> recommendations<br />

should sustain biotic diversity and natural<br />

processes by managing several <strong>for</strong>est and woodland<br />

systems used by <strong>the</strong> owl. Recommendations<br />

include management of mixed-conifer and pineoak<br />

<strong>for</strong>ests, and riparian areas. Ponderosa pine<br />

and spruce-fir <strong>for</strong>ests are also considered to a<br />

limited degree.<br />

Several potential threats to <strong>the</strong> owl were<br />

identified by examining <strong>the</strong> best available in<strong>for</strong>mation<br />

on <strong>the</strong> owl’s biology, and by evaluating<br />

ecological disturbance patterns and current<br />

conditions throughout <strong>the</strong> owl’s range. Primary<br />

threats include catastrophic fire, timber and<br />

fuelwood harvest, grazing, and recreation. The<br />

magnitude of a threat’s influence on <strong>the</strong> owl can<br />

vary according to temporal and spatial setting.<br />

For this reason, general recommendations were<br />

Volume I/Part III<br />

122<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

developed by habitat type which apply throughout<br />

<strong>the</strong> owl’s range, and are emphasized according<br />

to <strong>the</strong> magnitude of <strong>the</strong> threats within each<br />

RU. The recommendations were also designed<br />

to provide different levels of protection depending<br />

on <strong>the</strong> owl’s use of a particular habitat, <strong>the</strong><br />

nature of <strong>the</strong> threats, and management potential.<br />

Our intent was to offer <strong>the</strong> most specific recommendations<br />

that <strong>the</strong> best available in<strong>for</strong>mation<br />

would permit while allowing land managers<br />

flexibility <strong>for</strong> implementing <strong>the</strong> recommendations.<br />

Three areas of management are provided<br />

under <strong>the</strong> general recommendations: protected<br />

areas, restricted areas, and o<strong>the</strong>r <strong>for</strong>est and<br />

woodland types. Protected areas receive <strong>the</strong><br />

highest level of protection. <strong>Recovery</strong> plan<br />

guidelines take precedence over o<strong>the</strong>r management<br />

guidelines in protected areas. Guidelines<br />

<strong>for</strong> restricted areas are less specific and operate in<br />

conjunction with existing management guidelines.<br />

Specific guidelines are not proposed <strong>for</strong><br />

o<strong>the</strong>r <strong>for</strong>est and woodland types.<br />

Protected areas are all occupied nest or roost<br />

areas, all areas with slope >40% where timber<br />

harvest has not occurred in <strong>the</strong> past 20 years,<br />

and all legally administered reserved lands.<br />

Protection of owl nest and roost areas will be<br />

established by designating an area of protection<br />

around an activity center (PAC). This will<br />

require (1) inventory of spotted owls be<strong>for</strong>e<br />

planning any management activity that will alter<br />

stand structure; (2) delineating PAC areas of 243<br />

ha (600 ac) <strong>for</strong> all known <strong>Mexican</strong> spotted owl<br />

sites, including sites located prior to proposed<br />

management activities; and (3) light burning in<br />

PACs if considered necessary and prudent to<br />

reduce risk of catastrophic loss. Fur<strong>the</strong>r, a fire<br />

abatement program is proposed to allow treatment<br />

of small fuels within PACs and minimize<br />

probabilities of catastrophic fire. The purpose of<br />

PACs is to provide refugia habitat until it can be<br />

demonstrated reliably that owl habitat can be<br />

created through management. In addition,<br />

harvest of trees 40% where timber harvest<br />

has not occurred in <strong>the</strong> past 20 years. However,<br />

light burning and prescribed natural fire management<br />

is permitted. Prescribed natural fire is


also encouraged on reserved lands (e.g., wilderness,<br />

Research Natural Areas) where appropriate.<br />

We recommend that management activities<br />

be restricted on some lands outside of protected<br />

areas because patterns of owl use can be expected<br />

to change over time. The guidelines depend<br />

upon <strong>for</strong>est or woodland type. Silvicultural<br />

prescriptions should emphasize measures to place<br />

stand conditions on a trajectory to become owl<br />

habitat where appropriate. Stands that currently<br />

meet or exceed threshold conditions are subject<br />

to more stringent restrictions than o<strong>the</strong>r stands.<br />

Specific management prescriptions should be site<br />

specific and will vary according to short- or<br />

long-term objectives.<br />

Short-term guidelines should not be misconstrued<br />

as onetime management events. For<br />

example, large trees and snags are used by <strong>the</strong><br />

spotted owl and will continue to be needed<br />

beyond <strong>the</strong> life of <strong>the</strong> plan. Long-term guidelines<br />

are recommended <strong>for</strong> those activities and<br />

natural processes that combine to influence <strong>the</strong><br />

owl and its habitat beyond <strong>the</strong> life expectancy of<br />

this <strong>Recovery</strong> <strong>Plan</strong>.<br />

In addition, riparian communities should be<br />

managed by maintaining broad-leaved <strong>for</strong>ests in<br />

healthy condition where <strong>the</strong>y occur, especially in<br />

canyon-bottoms. Restoration may be necessary<br />

where such <strong>for</strong>ests are not regenerating adequately.<br />

Conceivably, restored riparian <strong>for</strong>ests<br />

could contribute additional nesting, wintering<br />

and dispersal habitat in <strong>the</strong> future. A mix of<br />

plant size and age classes should be emphasized<br />

in this community, to include large mature trees,<br />

vertical diversity, and o<strong>the</strong>r structural characteristics.<br />

No specific guidelines are recommended in<br />

<strong>for</strong>est or woodland types not typically used by<br />

<strong>the</strong> owl <strong>for</strong> nesting. These include ponderosa<br />

pine, spruce-fir, pinyon-juniper, and quaking<br />

aspen in areas outside of PACs. However, some<br />

relevant management of <strong>the</strong>se communities may<br />

produce desirable results <strong>for</strong> owl recovery.<br />

Examples of guidelines include managing <strong>for</strong><br />

landscape diversity, mimicking natural disturbance<br />

patterns, incorporating natural variation<br />

in stand conditions, retaining special features<br />

such as snags, and utilizing fire in an appropriate<br />

manner.<br />

Livestock and wildlife grazing may influence<br />

spotted owls by altering (1) prey availability, (2)<br />

fire risk of some habitats, (3) riparian plant<br />

communities, and (4) development of spotted<br />

Volume I/Part III<br />

123<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

owl habitat. The Team strongly advocates field<br />

monitoring and experimental research related to<br />

impacts of grazing on <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

O<strong>the</strong>r specific guidelines include (1) monitoring<br />

grazing use by livestock and key wildlife species<br />

(e.g. elk, deer), (2) implementing and en<strong>for</strong>cing<br />

grazing utilization standards that attain good to<br />

excellent range use standards, and (3) protecting<br />

or restoring riparian communities. These guidelines<br />

are emphasized in protected, restricted, and<br />

riparian areas.<br />

Several guidelines <strong>for</strong> managing recreation in<br />

protected, restricted, and riparian areas are<br />

recommended. These include: (1) no construction,<br />

ei<strong>the</strong>r of new facilities or <strong>for</strong> expanding<br />

existing facilities, is allowed within PACs during<br />

<strong>the</strong> breeding season; (2) construction during <strong>the</strong><br />

nonbreeding season should be considered on a<br />

case-specific basis; (3) managers should, on a<br />

case-specific basis, assess <strong>the</strong> presence and<br />

intensity of allowable recreational activities<br />

within PACs; and (4) seasonal closures of specifically<br />

designated recreation activities should be<br />

considered in extreme circumstances.<br />

Several important questions regarding <strong>the</strong><br />

owl’s ecology, and in particular about <strong>the</strong> effects<br />

of different management activities on <strong>the</strong> owl’s<br />

population viability, still remain. The Team<br />

recommends additional research on <strong>Mexican</strong><br />

spotted owl dispersal, genetics, habitat ecology,<br />

and population biology. Key in<strong>for</strong>mation that is<br />

vital <strong>for</strong> refining recovery strategies include (1)<br />

<strong>the</strong> degree of demographic and genetic isolation<br />

among subpopulations; (2) <strong>the</strong> relationship<br />

between fitness and specific habitat components;<br />

(3) population trend. Communication and<br />

collaboration between researchers and managers<br />

will be paramount <strong>for</strong> obtaining necessary<br />

in<strong>for</strong>mation.<br />

This <strong>Recovery</strong> <strong>Plan</strong> presents realistic goals<br />

<strong>for</strong> recovery of <strong>the</strong> <strong>Mexican</strong> spotted owl and its<br />

ultimate delisting. The goals are flexible in that<br />

<strong>the</strong>y allow local land managers to make sitespecific<br />

decisions about management <strong>for</strong> recovery.<br />

The success of <strong>the</strong> recovery process hinges<br />

on commitment and coordination among<br />

Federal and State land management agencies,<br />

sovereign Indian Nations, and <strong>the</strong> private sector<br />

to ensure that <strong>the</strong> plan is followed and executed<br />

as intended by <strong>the</strong> Team.


<strong>Recovery</strong> <strong>Plan</strong> Implementation<br />

Volume I, Part IV


Part IV discusses laws, regulations, policies,<br />

and authorities directly relevant to implementing<br />

<strong>the</strong> recovery recommendations included in Part<br />

III. An approach to implementation oversight is<br />

also recommended. Finally, a stepdown outline<br />

of recovery tasks and an implementation schedule<br />

are provided.<br />

This <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong> is<br />

based or predicated upon laws that designate<br />

specific legal authority and responsibility to<br />

government agencies <strong>for</strong> managing public<br />

resources, including wildlife and wildlife habitat.<br />

The following summarizes relevant laws and<br />

authorities applicable to implementation of this<br />

<strong>Recovery</strong> <strong>Plan</strong>.<br />

ENDANGERED ENDANGERED SPECIES SPECIES ACT ACT<br />

ACT<br />

Section 2(c)(2) of <strong>the</strong> Act expresses <strong>the</strong><br />

policy of Congress that “...all Federal departments<br />

and agencies shall seek to conserve endangered<br />

species and threatened species and shall<br />

utilize <strong>the</strong>ir authorities in fur<strong>the</strong>rance of <strong>the</strong><br />

purposes of [<strong>the</strong>] Act.” Section 7(a)(1) of <strong>the</strong><br />

Act requires Federal agencies to “...utilize <strong>the</strong>ir<br />

authorities in fur<strong>the</strong>rance of <strong>the</strong> purposes of <strong>the</strong><br />

Act by carrying out programs <strong>for</strong> <strong>the</strong> conservation<br />

of endangered species and threatened<br />

species....” Thus, Congress clearly intended<br />

conservation of endangered and threatened<br />

species to be considered in implementation of<br />

Federal programs and actions. In addition, o<strong>the</strong>r<br />

Federal laws and regulations require consideration<br />

of endangered and threatened species in<br />

program implementation, including <strong>the</strong> National<br />

Forest Management Act (NFMA) and <strong>the</strong><br />

National Environmental Policy Act (NEPA).<br />

Implementation of <strong>the</strong> Act is <strong>the</strong> responsibility<br />

of <strong>the</strong> Secretary of <strong>the</strong> Interior <strong>for</strong> listed<br />

terrestrial species. The Secretary generally delegates<br />

implementation authority to <strong>the</strong> FWS.<br />

The following sections of <strong>the</strong> Act are relevant to<br />

implementation of species recovery ef<strong>for</strong>ts:<br />

Volume I/Part IV<br />

A. A. A. IMPLEMENTING IMPLEMENTING LAWS, LAWS, LAWS, REGULATIONS,<br />

REGULATIONS,<br />

AND AND AND AUTHORITIES<br />

AUTHORITIES<br />

124<br />

Section Section 4<br />

4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Section 4 includes <strong>the</strong> listing and recovery<br />

provisions of <strong>the</strong> Act, which are discussed in<br />

detail in Part I. Section 4(b) of <strong>the</strong> Act provides<br />

<strong>for</strong> designation of critical habitat <strong>for</strong> endangered<br />

and threatened species. Regulations governing<br />

critical habitat designation are codified at 50<br />

CFR 424. Protection of critical habitat is administered<br />

under section 7 of <strong>the</strong> Act (discussed<br />

below). Critical habitat is defined under section<br />

3(5)(A) of <strong>the</strong> Act as:<br />

“(i) <strong>the</strong> specific areas within <strong>the</strong> geographical<br />

area occupied by <strong>the</strong> species...on<br />

which are found those physical or<br />

biological features (I) essential to <strong>the</strong><br />

conservation of <strong>the</strong> species and (II)<br />

which may require special management<br />

considerations or protection; and<br />

(ii) specific areas outside <strong>the</strong> geographical<br />

area occupied by <strong>the</strong> species...upon a<br />

determination by <strong>the</strong> Secretary that such<br />

areas are essential <strong>for</strong> <strong>the</strong> conservation of<br />

<strong>the</strong> species.”<br />

Section 4(d) of <strong>the</strong> Act provides <strong>for</strong> promulgation<br />

of special rules <strong>for</strong> threatened<br />

species only. This allows <strong>the</strong> Secretary to issue<br />

regulations as deemed necessary <strong>for</strong> <strong>the</strong> conservation<br />

of such species. Special rules can be useful<br />

in enacting regulatory provisions uniquely<br />

applicable to <strong>the</strong> species at hand, and can be<br />

promulgated to avoid unnecessary regulatory<br />

burden. For example, <strong>the</strong> FWS is considering a<br />

special 4(d) rule to allow small landowners in <strong>the</strong><br />

Pacific Northwest to harvest timber and conduct<br />

o<strong>the</strong>r activities without risk of violating <strong>the</strong><br />

prohibition of incidentally taking (see definition<br />

under Section 9, below) nor<strong>the</strong>rn spotted owls.


Section Section 5<br />

5<br />

Section 5 directs <strong>the</strong> Secretary to utilize<br />

funds and authorities of o<strong>the</strong>r laws in acquisition<br />

of lands, as deemed appropriate <strong>for</strong> conservation<br />

of endangered and threatened species.<br />

Section Section 6<br />

6<br />

This section authorizes cooperation with <strong>the</strong><br />

States in conservation of threatened and endangered<br />

species. Among its provisions is <strong>the</strong> authority<br />

to enter into management agreements<br />

and cooperative agreements and to allocate funds<br />

to <strong>the</strong> States that have entered into such agreements.<br />

Section Section 7<br />

7<br />

Section 7 and its implementing regulations<br />

at 50 CFR 402 govern cooperation between<br />

Federal agencies. Federal agencies must, in<br />

consultation with and with <strong>the</strong> assistance of <strong>the</strong><br />

Secretary, ensure that any action <strong>the</strong>y fund,<br />

authorize, or carry out is not likely to jeopardize<br />

<strong>the</strong> continued existence of listed species or result<br />

in <strong>the</strong> destruction or adverse modification of a<br />

listed species’ designated critical habitat. Regulations<br />

at 50 CFR 402 provide <strong>the</strong> following<br />

definitions:<br />

“‘Jeopardize <strong>the</strong> continued existence of’<br />

means to engage in an action that reasonably<br />

would be expected, directly or indirectly,<br />

to reduce appreciably <strong>the</strong> likelihood<br />

of both survival and recovery of a listed<br />

species in <strong>the</strong> wild by reducing <strong>the</strong> reproduction,<br />

numbers, or distribution of that<br />

species.”<br />

“‘Destruction or adverse modification’<br />

means a direct or indirect alteration that<br />

appreciably diminishes <strong>the</strong> value of critical<br />

habitat <strong>for</strong> both <strong>the</strong> survival and recovery<br />

of a listed species.”<br />

Section 7 requires action agencies to assess<br />

<strong>the</strong> effects of proposed actions on listed species<br />

and <strong>the</strong>ir critical habitat. If, as a result of that<br />

Volume I/Part IV<br />

125<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

assessment, <strong>the</strong> agency determines that an action<br />

may affect a listed species or its critical habitat,<br />

<strong>the</strong> agency must enter into consultation with <strong>the</strong><br />

FWS. That consultation may result in a biological<br />

opinion from <strong>the</strong> FWS, in which a determination<br />

is made as to whe<strong>the</strong>r jeopardy to <strong>the</strong><br />

species and/or destruction or adverse modification<br />

of its critical habitat are likely to result from<br />

<strong>the</strong> agency action.<br />

If a biological opinion concludes that jeopardy<br />

to <strong>the</strong> species and/or adverse modification<br />

of its critical habitat are not likely to result from<br />

a proposed action, <strong>the</strong> action may proceed. The<br />

FWS may provide conservation recommendations<br />

to <strong>the</strong> agency on ways to minimize or<br />

avoid potential adverse effects on listed species<br />

and/or critical habitat. Implementation of <strong>the</strong>se<br />

conservation recommendations is at <strong>the</strong> action<br />

agencies’ discretion. In cases where <strong>the</strong> action is<br />

likely to result in <strong>the</strong> incidental taking of a<br />

species (see definition under “Section 9,” below),<br />

<strong>the</strong> Service may provide reasonable and prudent<br />

measures to minimize <strong>the</strong> amount or extent of<br />

incidental take. The terms and conditions that<br />

accompany and implement any reasonable and<br />

prudent measures are nondiscretionary and must<br />

be implemented. However, reasonable and<br />

prudent measures and <strong>the</strong>ir implementing terms<br />

and conditions cannot alter <strong>the</strong> basic design,<br />

location, scope, duration, or timing of <strong>the</strong><br />

action; and <strong>the</strong>y may involve only minor<br />

changes.<br />

If a biological opinion determines that<br />

jeopardy and/or adverse modification is likely to<br />

result from a proposed action, <strong>the</strong> FWS and <strong>the</strong><br />

action agency develop reasonable and prudent<br />

alternatives, if any, to <strong>the</strong> proposed action.<br />

Reasonable and prudent alternatives refer to<br />

alternative actions that are consistent with <strong>the</strong><br />

intended purpose of <strong>the</strong> proposed action, that<br />

can be implemented within <strong>the</strong> action agency’s<br />

legal authority, that are economically and technologically<br />

feasible, and that <strong>the</strong> FWS believes<br />

will not result in jeopardy to listed species or<br />

destruction or adverse modification of critical<br />

habitat. If no reasonable or prudent alternatives<br />

can be identified, <strong>the</strong> action agency may apply to<br />

<strong>the</strong> Endangered Species Committee <strong>for</strong> an<br />

exemption to <strong>the</strong> prohibition of jeopardy and/or


destruction or adverse modification of critical<br />

habitat.<br />

Section Section 8<br />

8<br />

Section 8 authorizes international cooperation<br />

in conservation of endangered and threatened<br />

species. Included under this section is <strong>the</strong><br />

authority to provide financial assistance to<br />

<strong>for</strong>eign countries to assist in <strong>the</strong>ir conservation<br />

ef<strong>for</strong>ts.<br />

Section Section 9<br />

9<br />

Section 9 covers prohibited acts in regard to<br />

listed species. Of relevance to <strong>the</strong> <strong>Mexican</strong><br />

spotted owl is <strong>the</strong> prohibition of taking individuals.<br />

“Take” is defined as “...to harass, harm,<br />

pursue, shoot, wound, kill, trap, capture, or<br />

collect, or to attempt to engage in any such<br />

conduct.” Permits <strong>for</strong> direct taking of threatened<br />

species may be issued <strong>for</strong> scientific purposes,<br />

to enhance propagation or survival, in<br />

cases of economic hardship, <strong>for</strong> zoological<br />

exhibition, or <strong>for</strong> educational purposes (50 CFR<br />

17.32).<br />

Taking of spotted owls is most likely to<br />

occur through “incidental take.” “Incidental<br />

take” is defined as taking that results from, but is<br />

not <strong>the</strong> purpose of, carrying out an o<strong>the</strong>rwise<br />

lawful activity. Incidental taking of spotted owls<br />

may result from such activities as timber harvest,<br />

if that activity results in habitat loss to an extent<br />

that an individual spotted owl’s normal behavior<br />

patterns are impaired. In cases where incidental<br />

taking will not result in jeopardy to a listed<br />

species, <strong>the</strong> FWS may issue an incidental take<br />

statement in a biological opinion on a proposed<br />

Federal action, <strong>the</strong>reby removing <strong>the</strong> take<br />

prohibition. Relief from <strong>the</strong> taking prohibition<br />

<strong>for</strong> non-Federal activities is discussed under<br />

“section 10” below.<br />

Section Section 10<br />

10<br />

Section 10 authorizes <strong>the</strong> FWS to issue<br />

permits <strong>for</strong> takings o<strong>the</strong>rwise prohibited under<br />

section 9. Such permits may be issued <strong>for</strong> research<br />

purposes and <strong>the</strong> o<strong>the</strong>r situations de-<br />

Volume I/Part IV<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

scribed above. In addition, section 10(a)(1)(B)<br />

allows permits <strong>for</strong> incidental taking that may<br />

result from an activity, provided an applicant<br />

submits a conservation plan that specifies:<br />

“(i) <strong>the</strong> impact which will likely result<br />

from such taking;<br />

(ii) what steps <strong>the</strong> applicant will take to<br />

minimize and mitigate such impacts,<br />

and <strong>the</strong> funding that will be available<br />

to implement such steps;<br />

(iii) what alternative actions to such taking<br />

<strong>the</strong> applicant considered and <strong>the</strong><br />

reasons why such alternatives are not<br />

being utilized; and<br />

(iv) such o<strong>the</strong>r measures that <strong>the</strong> [FWS]<br />

may require as being necessary or<br />

appropriate <strong>for</strong> purposes of <strong>the</strong> plan.”<br />

NATIONAL NATIONAL FOREST<br />

FOREST<br />

MANAGEMENT MANAGEMENT ACT<br />

ACT<br />

The NFMA governs Forest Service Management<br />

on National Forest System lands. Section<br />

219.19 (Fish and wildlife resources) states:<br />

“Fish and wildlife habitat shall be managed<br />

to maintain viable populations of existing<br />

native and desired nonnative vertebrate<br />

species in <strong>the</strong> planning area. For<br />

planning purposes, a viable population<br />

shall be regarded as one which has <strong>the</strong><br />

estimated numbers and distribution of<br />

reproductive individuals to ensure its<br />

continued existence is well distributed in<br />

<strong>the</strong> planning area. In order to ensure that<br />

viable populations will be maintained,<br />

habitat must be provided to support, at<br />

least, a minimum number of reproductive<br />

individuals and that habitat must be<br />

well distributed so that those individuals<br />

can interact with o<strong>the</strong>rs in <strong>the</strong> planning<br />

area.”<br />

In <strong>for</strong>mulating alternatives during project<br />

planning, <strong>the</strong> following is required in regard to<br />

fish and wildlife habitat:<br />

126


“Each alternative shall establish objectives <strong>for</strong><br />

<strong>the</strong> maintenance and improvement of<br />

habitat <strong>for</strong> management indicator species...to<br />

<strong>the</strong> degree consistent with overall multipleuse<br />

objectives of <strong>the</strong> alternative. To meet this<br />

goal, management planning <strong>for</strong> <strong>the</strong> fish and<br />

wildlife resources shall meet <strong>the</strong> requirements<br />

set <strong>for</strong>th [as follows:]<br />

(1) In order to estimate <strong>the</strong> effects of each<br />

alternative on fish and wildlife populations,<br />

certain vertebrate and/or invertebrate<br />

species present in <strong>the</strong> area shall be<br />

identified and selected as management<br />

indicator species and <strong>the</strong> reasons <strong>for</strong> <strong>the</strong>ir<br />

selection will be stated. These species<br />

shall be selected because <strong>the</strong>ir population<br />

changes are believed to indicate <strong>the</strong><br />

effects of management activities. In <strong>the</strong><br />

selection of management indicator<br />

species, <strong>the</strong> following categories shall be<br />

represented where appropriate:<br />

Volume I/Part IV<br />

* Endangered and threatened plant and<br />

animal species identified on State and<br />

Federal lists <strong>for</strong> <strong>the</strong> planning area;<br />

* Species with special habitat needs that<br />

may be influenced significantly by<br />

planned management programs;<br />

* Species commonly hunted, fished, or<br />

trapped;<br />

* Nongame species of special interest;<br />

and<br />

* Additional plant or animal species<br />

selected because <strong>the</strong>ir population<br />

changes are believed to indicate <strong>the</strong><br />

effects of management activities on<br />

o<strong>the</strong>r species of selected major biological<br />

communities or on water quality.<br />

“On <strong>the</strong> basis of available scientific<br />

in<strong>for</strong>mation, <strong>the</strong> interdisciplinary team<br />

shall estimate <strong>the</strong> effects of changes in<br />

vegetation type, timber age classes,<br />

community composition, rotation age,<br />

and year-long suitability of habitat<br />

127<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

related to mobility of management<br />

indicator species. Where appropriate,<br />

measures to mitigate adverse effects shall<br />

be prescribed.<br />

(2) <strong>Plan</strong>ning alternatives shall be stated and<br />

evaluated in terms of both amount and<br />

quality of habitat and of animal population<br />

trends of <strong>the</strong> management indicator<br />

species.<br />

(3) Biologists from State fish and wildlife<br />

agencies and o<strong>the</strong>r Federal agencies shall<br />

be consulted in order to coordinate<br />

planning <strong>for</strong> fish and wildlife, including<br />

opportunities <strong>for</strong> <strong>the</strong> reintroduction of<br />

extirpated species.<br />

(4) Access and dispersal problems of hunting,<br />

fishing, and o<strong>the</strong>r visitor uses shall<br />

be considered.<br />

(5) The effects of pest and fire management<br />

on fish and wildlife populations shall be<br />

considered.<br />

(6) Population trends of <strong>the</strong> management<br />

indicator species will be monitored and<br />

relationships to habitat changes determined.<br />

This monitoring will be done in<br />

cooperation with State fish and wildlife<br />

agencies, to <strong>the</strong> extent practicable.<br />

(7) Habitat determined to be critical <strong>for</strong><br />

threatened and endangered species shall<br />

be identified, and measures shall be<br />

prescribed to prevent <strong>the</strong> destruction or<br />

adverse modification of such habitat.<br />

Objectives shall be determined <strong>for</strong><br />

threatened and endangered species that<br />

shall provide <strong>for</strong>, where possible, <strong>the</strong>ir<br />

removal from listing as threatened and<br />

endangered species through appropriate<br />

conservation measures, including <strong>the</strong><br />

designation of special areas to meet <strong>the</strong><br />

protection and management needs of<br />

such species.”


NATIONAL NATIONAL ENVIRONMENTAL<br />

ENVIRONMENTAL<br />

POLICY POLICY ACT<br />

ACT<br />

The NEPA requires Federal agencies to<br />

prepare Environmental Impact Statements (EIS)<br />

or Environmental Assessments (EA) <strong>for</strong> implementation<br />

of agency actions and issuance or<br />

modification of agency policies and guidance.<br />

Impacts of <strong>the</strong> proposed action or policy amendment<br />

on endangered and threatened species<br />

must be evaluated. If a deciding official determines<br />

that no significant impact will result from<br />

an action or policy amendment, a Finding of No<br />

Significant Impact is issued. If an agency determines<br />

that a significant impact will result from<br />

<strong>the</strong> proposed action or policy amendment, an<br />

EIS must be prepared. An EIS addresses a range<br />

of alternatives. It is released <strong>for</strong> public review<br />

and comment, after which an alternative is<br />

selected and a Record of Decision is signed by<br />

<strong>the</strong> deciding official.<br />

MIGRATORY MIGRATORY MIGRATORY BIRD BIRD TREATY TREATY ACT<br />

ACT<br />

Prior to listing <strong>the</strong> <strong>Mexican</strong> spotted owl as<br />

threatened, <strong>the</strong> Migratory Bird Treaty Act<br />

(MBTA) provided <strong>the</strong> only Federal protection<br />

<strong>for</strong> <strong>the</strong> subspecies o<strong>the</strong>r than that af<strong>for</strong>ded by<br />

land-management agencies. Under <strong>the</strong> provisions<br />

of <strong>the</strong> MBTA, it is unlawful to pursue,<br />

hunt, take, capture, or kill in any manner any<br />

migratory bird unless permitted by regulations.<br />

The MBTA applies in both <strong>the</strong> U.S. and<br />

Mexico. Because <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

exhibits migratory behavior in some areas it is<br />

included on <strong>the</strong> list of birds protected under <strong>the</strong><br />

MBTA.<br />

Volume I/Part IV<br />

TRIBAL TRIBAL LANDS<br />

LANDS<br />

The <strong>Recovery</strong> Team encourages adoption of<br />

<strong>the</strong> recovery recommendations by all Tribes<br />

administering lands that support <strong>Mexican</strong><br />

spotted owl habitat. Tribal land-management<br />

regulations and programs, including those <strong>for</strong><br />

conservation of species, typically require enactment<br />

by Tribal Councils.<br />

128<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

STATE STATE AND AND PRIVATE PRIVATE LANDS LANDS<br />

LANDS<br />

Although relatively few <strong>Mexican</strong> spotted<br />

owls are known on State and private lands, <strong>the</strong><br />

Team recommends that States continue and/or<br />

begin a program to inventory <strong>for</strong>ested areas <strong>for</strong><br />

<strong>the</strong> presence of <strong>Mexican</strong> spotted owls. The<br />

<strong>Recovery</strong> Team is unaware of any State laws or<br />

regulations that govern management of spotted<br />

owl habitat on State or private lands. The <strong>Recovery</strong><br />

Team recommends incorporating <strong>the</strong> recovery<br />

recommendations into State wildlife and<br />

<strong>for</strong>est practices laws and regulations. In addition,<br />

<strong>the</strong> <strong>Recovery</strong> Team encourages <strong>the</strong> FWS to<br />

evaluate <strong>the</strong> importance of State and private<br />

lands to <strong>the</strong> <strong>Mexican</strong> spotted owl, and to consider<br />

promulgating a special rule under section<br />

4(d) of <strong>the</strong> Act that specifies habitat-altering<br />

activities that can be allowed on private lands<br />

without violating <strong>the</strong> prohibition of incidentally<br />

taking <strong>Mexican</strong> spotted owls.<br />

MEXICO MEXICO<br />

MEXICO<br />

The <strong>Recovery</strong> Team is unfamiliar with <strong>the</strong><br />

laws, regulations, and authorities that are available<br />

or appropriate <strong>for</strong> implementing <strong>the</strong> recovery<br />

recommendations in Mexico. As recommended<br />

later in Part IV, <strong>the</strong> <strong>Recovery</strong> Team<br />

expects <strong>the</strong> FWS to arrange a meeting with<br />

<strong>Mexican</strong> officials to discuss <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

and its implementation.


Volume I/Part IV<br />

RECOVERY RECOVERY UNIT<br />

UNIT<br />

WORKING WORKING TEAMS TEAMS<br />

TEAMS<br />

The Team strongly recommends <strong>for</strong>mation<br />

of interagency working teams whose responsibility<br />

would be to oversee <strong>the</strong> implementation of<br />

<strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>. These <strong>Recovery</strong> Unit Working<br />

Teams would coordinate with and report to<br />

<strong>the</strong> <strong>Recovery</strong> Team, which would evaluate any<br />

Working Team recommendations be<strong>for</strong>e passing<br />

<strong>the</strong>m on to <strong>the</strong> FWS. Working Teams <strong>for</strong> each<br />

U.S. <strong>Recovery</strong> Unit should be appointed by <strong>the</strong><br />

FWS as subunits under <strong>the</strong> <strong>Recovery</strong> Team<br />

umbrella. <strong>Recovery</strong> Team members may also<br />

serve on <strong>Recovery</strong> Unit Working Teams if that<br />

arrangement is agreeable. Membership of <strong>the</strong><br />

Working Teams should include, at a minimum,<br />

one representative from each of <strong>the</strong> following:<br />

1. Each involved FWS Ecological Services<br />

Field Office<br />

2. Each involved FS Region<br />

3. Each involved State<br />

4. Each involved Indian Reservation<br />

5. Any o<strong>the</strong>r involved agency (e.g., BLM,<br />

NPS).<br />

Each Working Team should have a research<br />

scientist among its membership. That person<br />

may be affiliated with one of <strong>the</strong> agencies listed<br />

above, or may be independent. In addition to<br />

<strong>the</strong> above, o<strong>the</strong>r interested persons approved by<br />

<strong>the</strong> <strong>Recovery</strong> Team and <strong>the</strong> FWS should be<br />

allowed to participate if <strong>the</strong>y so request. Such<br />

participants may include a representative from a<br />

conservation organization, a representative from<br />

<strong>the</strong> timber or o<strong>the</strong>r affected industry, a representative<br />

from an interested county or o<strong>the</strong>r local<br />

government agency, and o<strong>the</strong>rs as appropriate.<br />

Such a diverse membership would allow ideas of<br />

varying viewpoints to be discussed and would<br />

allow local interested parties to participate in<br />

B. B. B. IMPLEMENTATION<br />

IMPLEMENTATION<br />

OVERSIGHT<br />

OVERSIGHT<br />

129<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

plan implementation and resolution of local<br />

issues. Working Teams <strong>for</strong> each <strong>Mexican</strong> <strong>Recovery</strong><br />

unit should be similarly composed.<br />

Once <strong>the</strong> FWS <strong>for</strong>mulates a membership<br />

list, that list should be submitted to <strong>the</strong> <strong>Recovery</strong><br />

Team <strong>for</strong> review. The <strong>Recovery</strong> Team would<br />

<strong>the</strong>n request <strong>the</strong> FWS’s Southwest Regional<br />

Director’s approval. Travel costs <strong>for</strong> each member<br />

would be borne by <strong>the</strong> member’s agency or<br />

organization.<br />

The functions of <strong>the</strong> <strong>Recovery</strong> Unit Working<br />

Teams should include <strong>the</strong> following:<br />

1. Provide technical assistance to agencies<br />

and landowners on such issues as project<br />

designs, spotted owl management plan<br />

development, and <strong>Recovery</strong> <strong>Plan</strong> compliance.<br />

The <strong>Recovery</strong> Team strongly<br />

encourages conducting <strong>Recovery</strong> <strong>Plan</strong><br />

implementation workshops to provide<br />

biologists, <strong>for</strong>esters, and o<strong>the</strong>r landmanagement<br />

personnel a common<br />

working knowledge of <strong>the</strong> provisions of<br />

this <strong>Recovery</strong> <strong>Plan</strong>. For example, a workshop<br />

to develop procedures <strong>for</strong> delineating<br />

PACs would encourage consistent<br />

application of recovery recommendations.<br />

Specific workshop recommendations<br />

are provided in IV.C.<br />

2. Provide guidance and interpretation on<br />

implementation of <strong>the</strong> recommendations<br />

contained in this <strong>Recovery</strong> <strong>Plan</strong>.<br />

3. Provide research assistance by procuring<br />

financial and logistic support, screening<br />

research proposals <strong>for</strong> importance and<br />

relevance, recommending to <strong>the</strong> <strong>Recovery</strong><br />

Team prioritization of research<br />

proposals, and o<strong>the</strong>r functions.<br />

4. Recommend <strong>Recovery</strong> <strong>Plan</strong> revisions<br />

based on research results that may<br />

enhance recovery ef<strong>for</strong>ts in that<br />

specific RU.


5. Prioritize areas to be inventoried within<br />

<strong>the</strong> RU.<br />

6. Promote communication between<br />

various local interests and help resolve<br />

conflicting interpretations of <strong>the</strong> <strong>Recovery</strong><br />

<strong>Plan</strong> provisions.<br />

7. Monitor plan implementation and<br />

report problems, successes, and general<br />

recovery progress to <strong>the</strong> FWS and <strong>the</strong><br />

<strong>Recovery</strong> Team at least annually.<br />

CONTINUING CONTINUING DUTIES DUTIES OF OF THE<br />

THE<br />

RECOVERY RECOVERY TEAM<br />

TEAM<br />

The <strong>Recovery</strong> Team recommends that it be<br />

continued throughout <strong>Recovery</strong> <strong>Plan</strong> implementation.<br />

Once <strong>the</strong> final <strong>Recovery</strong> <strong>Plan</strong> is complete,<br />

<strong>the</strong> <strong>Recovery</strong> Team should meet at least<br />

twice per year <strong>for</strong> <strong>the</strong> first two years and annually<br />

<strong>the</strong>reafter. The purpose of <strong>the</strong>se meetings<br />

would be to hear and discuss plan implementation<br />

reports with <strong>the</strong> <strong>Recovery</strong> Unit Working<br />

Teams, and to report to <strong>the</strong> FWS on <strong>the</strong> progress<br />

of <strong>the</strong> recovery ef<strong>for</strong>t. The Team would also<br />

consider recommendations from <strong>Recovery</strong> Unit<br />

Working Teams and decide what recommendations<br />

should be brought <strong>for</strong>ward to <strong>the</strong> FWS as<br />

potential revisions to <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>.<br />

Volume I/Part IV<br />

130<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

CENTRALIZED CENTRALIZED SPOTTED SPOTTED OWL<br />

OWL<br />

INFORMATION INFORMATION REPOSITORY<br />

REPOSITORY<br />

The <strong>Recovery</strong> Team recommends that a<br />

central <strong>Mexican</strong> spotted owl data facility be<br />

maintained throughout <strong>the</strong> life of <strong>the</strong> <strong>Recovery</strong><br />

<strong>Plan</strong>. The main purpose of such a facility would<br />

be to house a spotted owl GIS database, including<br />

data assembled through <strong>the</strong> monitoring<br />

program, inventory program, and o<strong>the</strong>r programs<br />

recommended in this <strong>Recovery</strong> <strong>Plan</strong>. In<br />

addition, <strong>the</strong> facility would maintain and periodically<br />

update a <strong>Mexican</strong> spotted owl bibliography.<br />

Such a facility would be a valuable resource<br />

<strong>for</strong> biologists, land managers, researchers, and<br />

o<strong>the</strong>rs who may need in<strong>for</strong>mation throughout<br />

<strong>the</strong> plan implementation period. Considerable<br />

in<strong>for</strong>mation, assembled as a result of development<br />

of this plan, is already stored in a GIS<br />

system maintained by <strong>the</strong> National Biological<br />

Service’s Midcontinent Ecological Science<br />

Center (<strong>for</strong>merly <strong>the</strong> National Ecology Research<br />

Center) in Fort Collins, Colorado; continuance<br />

of that arrangement is recommended by <strong>the</strong><br />

Team. In addition, a considerable “Literature<br />

Cited” section is included in this plan, which<br />

should provide a good start to development of a<br />

<strong>Mexican</strong> spotted owl bibliography.


This section lists specific tasks that need to<br />

be implemented according to <strong>the</strong> recovery<br />

recommendations in Part III, plus <strong>Recovery</strong> <strong>Plan</strong><br />

oversight provisions discussed earlier in Part IV.<br />

This list is in a stepdown <strong>for</strong>mat, as required in<br />

<strong>the</strong> FWS recovery planning guidelines. Each task<br />

is also listed in Table IV.D.1, where <strong>the</strong> responsible<br />

parties <strong>for</strong> task implementation and <strong>the</strong><br />

estimated costs of carrying out <strong>the</strong> tasks are<br />

provided. Tasks are categorized as follows:<br />

1. Resource Management Programs. Many of<br />

<strong>the</strong> recovery recommendations relate to<br />

spotted owl considerations that shouldbe<br />

incorporated into planning <strong>for</strong> o<strong>the</strong>r<br />

resource management objectives such as<br />

timber harvest, recreation, and management<br />

of o<strong>the</strong>r species.<br />

2. Active Management. These recovery<br />

tasks are to be implemented actively.<br />

They include <strong>for</strong>est health enhancement<br />

and protection, riparian restoration, and<br />

development of a long-term spotted owl<br />

management plan.<br />

3. Monitoring. These recommendations<br />

relate to monitoring <strong>the</strong> spotted owl<br />

population and habitat.<br />

4. Research. These recommendations<br />

include research studies designed to<br />

increase life-history knowledge of <strong>the</strong><br />

subspecies and to test <strong>the</strong> effects of land<br />

management activities on spotted owls.<br />

5. Oversight, Review, Evaluation, and<br />

Revision. These tasks are necessary to<br />

monitor <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>’s effectiveness<br />

and to determine if and when <strong>Recovery</strong><br />

<strong>Plan</strong> revision is necessary.<br />

Volume I/Part IV<br />

C. C. STEPDOWN STEPDOWN STEPDOWN OUTLINE<br />

OUTLINE<br />

131<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

1. 1. Resour esour esource esour ce M MManagement<br />

M anagement P PPrograms<br />

PP<br />

ograms<br />

11. Incorporate recovery recommendations<br />

(Part III) into land management<br />

programs.<br />

111. Conduct <strong>the</strong> NEPA process to<br />

amend appropriate land management<br />

guidance and policy documents<br />

(Federal lands).<br />

1111. FS<br />

1112. BLM<br />

1113. NPS<br />

1114. DOD<br />

112. Incorporate recovery recommendations<br />

into Tribal management plans.<br />

1121. White Mountain Apache<br />

1122. Mescalero Apache<br />

1123. San Carlos Apache<br />

1124. Navajo<br />

1125. O<strong>the</strong>r tribes<br />

113. Incorporate recovery recommendations<br />

into State regulations pertaining<br />

to timber harvests and o<strong>the</strong>r<br />

activities on State and private lands.<br />

1131. Arizona<br />

1132. New Mexico<br />

1133. Utah<br />

1134. Colorado<br />

114. Incorporate recovery recommendations<br />

into <strong>Mexican</strong> policy<br />

documents.<br />

1141. Arrange a meeting between<br />

FWS, <strong>Recovery</strong> Team, and<br />

<strong>Mexican</strong> representatives to<br />

discuss provisions of <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong>.


Volume I/Part IV<br />

1142. Conduct actions necessary<br />

to officially adopt <strong>Recovery</strong><br />

<strong>Plan</strong> recommendations into<br />

<strong>Mexican</strong> law and/or policy,<br />

as appropriate.<br />

12. Conduct pre–project <strong>Mexican</strong> spotted owl<br />

inventories in project areas.<br />

121. Federal agencies<br />

1211. FS<br />

1212. BLM<br />

1213. NPS<br />

1214. DOD<br />

122. Tribes<br />

1221. White Mountain Apache<br />

1222. Mescalero Apache<br />

1223. San Carlos Apache<br />

1224. Navajo<br />

1225. O<strong>the</strong>r tribes<br />

123. States<br />

1231. Arizona<br />

1232. New Mexico<br />

1233. Utah<br />

1234. Colorado<br />

124. Mexico<br />

2. 2. Activ ctiv ctive ctiv e M MManagement<br />

M anagement<br />

21. Develop and/or implement <strong>for</strong>est<br />

health improvement and protection<br />

programs.<br />

211. Federal lands<br />

2111. FS<br />

2112. BLM<br />

2113. NPS<br />

2114. DOD<br />

2115. O<strong>the</strong>r Federal agencies<br />

212. Tribal lands<br />

2121. White Mountain Apache<br />

2122. Mescalero Apache<br />

132<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

2123. San Carlos Apache<br />

2124. Navajo<br />

2125. O<strong>the</strong>r tribes<br />

213. State and private lands<br />

2131. Arizona<br />

2132. New Mexico<br />

2133. Utah<br />

2134. Colorado<br />

214. Mexico<br />

22. Actively manage riparian habitat<br />

(e.g., restore degraded areas).<br />

221. Lowland riparian<br />

2211. BLM<br />

2212. State of Arizona<br />

2213. State of New Mexico<br />

2214. State of Utah<br />

2215. State of Colorado<br />

2216. Mexico<br />

222. Middle to upper elevation riparian<br />

2221. Federal lands<br />

2222. Tribes<br />

22211. FS<br />

22212. BLM<br />

22213. NPS<br />

22214. DOD<br />

22215. O<strong>the</strong>r Federal<br />

agencies<br />

22221. White Mountain<br />

Apache<br />

22222. Mescalero Apache<br />

22223. San Carlos Apache<br />

22224. Navajo<br />

22225. O<strong>the</strong>r tribes<br />

2223. Mexico


23. Develop and implement a long–term,<br />

range wide management plan.<br />

Volume I/Part IV<br />

231. Establish and support a Federal/<br />

Tribal/State/<strong>Mexican</strong> team to<br />

develop <strong>the</strong> plan.<br />

232. Develop a draft management plan.<br />

233. Conduct peer/public review.<br />

234. Produce final management plan.<br />

235. Develop appropriate implementation<br />

documents.<br />

3. 3. M MMonitor<br />

M onitor onitoring onitor ing<br />

2351. Joint Federal agency EIS<br />

2352. White Mountain Apache<br />

2353. Mescalero Apache<br />

2354. San Carlos Apache<br />

2355. Navajo<br />

2356. O<strong>the</strong>r tribes<br />

2357. State MOUs<br />

23571. Arizona<br />

23572. New Mexico<br />

23573. Utah<br />

23574. Colorado<br />

2358. Mexico<br />

31. Implement <strong>the</strong> population monitoring<br />

program detailed in Part III.<br />

311. Secure funding <strong>for</strong> <strong>the</strong> entire<br />

monitoring period (up to 15<br />

years).<br />

312. Appoint a principle investigator.<br />

313. Develop detailed study methodology/protocols.<br />

314. Conduct <strong>Recovery</strong> Team/peer<br />

review of program.<br />

315. Conduct a pilot study.<br />

316. Evaluate and revise methodology/<br />

protocols.<br />

317. Implement <strong>the</strong> monitoring<br />

program.<br />

32. Implement <strong>the</strong> habitat monitoring<br />

program detailed in Part III.<br />

321. Macrohabitat<br />

133<br />

4. 4. Research<br />

Research<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

3211. Acquire appropriate remote<br />

sensing imagery.<br />

3212. Conduct necessary groundtruthing,<br />

imagery classification,<br />

geo-referencing, etc.<br />

3213. Acquire remote sensing<br />

imagery at year 5.<br />

3214. Conduct necessary groundtruthing,<br />

imager classification,<br />

geo-referencing, etc.<br />

3215. Conduct change-detection<br />

analysis.<br />

3216. Acquire remote sensing<br />

imagery at year 10.<br />

3217. Conduct necessary groundtruthing,<br />

imagery classification,<br />

geo-referencing, etc.<br />

3218. Conduct change-detection<br />

analysis.<br />

322. Microhabitat (ongoing)<br />

3221. Take pre-treatment measurements<br />

of relevant<br />

habitat variables.<br />

3222. Design treatment(s) to<br />

accomplish spotted owl<br />

habitat or o<strong>the</strong>r ecosystem<br />

management goals.<br />

3223. Conduct treatment<br />

3224. Take post-treatment measurements<br />

at year 1 of<br />

important habitat variables.<br />

3225. Compare pre- and posttreatment<br />

data to determine<br />

whe<strong>the</strong>r objectives of<br />

treatment were met.<br />

3226. Measure habitat variables at<br />

year 5.<br />

3227. Determine whe<strong>the</strong>r treated<br />

stands are on appropriate<br />

trajectories.<br />

41. Implement <strong>the</strong> research recommendations<br />

outlined in Part III.<br />

411. Conduct dispersal studies.


Volume I/Part IV<br />

4111. Examine connectivity of<br />

subpopulations within and<br />

between RUs.<br />

4112. Determine habitat configurations<br />

that best facilitate<br />

dispersal and enhance<br />

survival rates of dispersing<br />

juveniles.<br />

412. Conduct studies on genetics.<br />

4121. Determine whe<strong>the</strong>r and to<br />

what degree subpopulations<br />

are genetically isolated.<br />

4122. Determine <strong>the</strong> extent and<br />

patterns of gene flow across<br />

<strong>the</strong> landscape.<br />

413. Conduct habitat studies.<br />

4131. Study <strong>the</strong> extent to which<br />

habitat use is influenced<br />

by prey availability,<br />

microclimatic factors, or<br />

presence of predators.<br />

4132. Determine which habitat<br />

components influence<br />

individual fitness and<br />

population persistence.<br />

414. Study <strong>the</strong> effects of land-use practices<br />

on spotted owls and/or spotted<br />

owl habitat.<br />

4141. Determine <strong>the</strong> effects of<br />

various silvicultural and<br />

timber–harvest practices on<br />

spotted owl habitat.<br />

4142. Determine <strong>the</strong> effects of<br />

livestock and wildlife<br />

grazing on spotted owl<br />

habitat and prey.<br />

4143. Determine <strong>the</strong> effects of<br />

prescribed fire on spotted<br />

owl habitat and prey.<br />

4144. Determine <strong>the</strong> effects of<br />

recreational activities on<br />

spotted owl habitat.<br />

134<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

415. Study <strong>the</strong> effects of human<br />

disturbance on spotted owls.<br />

4151. Determine <strong>the</strong> effects of<br />

noise-producing activities<br />

on nesting spotted owls.<br />

4152. Determine <strong>the</strong> effects of<br />

suburban and rural development<br />

on habitats and<br />

populations of spotted owls.<br />

416. Study <strong>the</strong> effects of <strong>Recovery</strong> <strong>Plan</strong><br />

implementation on o<strong>the</strong>r ecosystem<br />

components.<br />

4161. Vertebrates and vertebrate<br />

communities<br />

4162. Invertebrates and<br />

invertebrate communities<br />

4163. <strong>Plan</strong>ts and plant<br />

communities<br />

4164. Abiotic features<br />

(e.g. hydrological systems)<br />

4165. Ecosystem structure and<br />

functioning<br />

42. Conduct general inventories in areas that<br />

have not previously been inventoried <strong>for</strong><br />

spotted owls.<br />

421. Federal lands<br />

4211. FS<br />

4212. BLM<br />

4213. NPS<br />

4214. DOD<br />

4215. O<strong>the</strong>r Federal agencies<br />

422. Tribal lands<br />

4221. White Mountain Apache<br />

4222. Mescalero Apache<br />

4223. San Carlos Apache<br />

4224. Navajo<br />

4225. O<strong>the</strong>r tribes<br />

423. State and private lands<br />

4231. Arizona<br />

4232. New Mexico


Volume I/Part IV<br />

4233. Utah<br />

4234. Colorado<br />

424. Mexico<br />

43. Maintain a centralized <strong>Mexican</strong> spotted<br />

owl in<strong>for</strong>mation facility.<br />

431. Establish and maintain a <strong>Mexican</strong><br />

spotted owl GIS database.<br />

4311. Establish and annually<br />

update spotted owl location<br />

records and inventory<br />

coverages.<br />

4312. Establish and periodically<br />

update spotted owl habitat<br />

coverages at varying spatial<br />

scales.<br />

432. Develop and periodically update a<br />

spotted owl bibliography.<br />

433. Distribute in<strong>for</strong>mation to land<br />

mangers and o<strong>the</strong>rs who request it.<br />

5. 5. Oversight, Oversight, Review, Review, Evaluation,<br />

Evaluation,<br />

and and Revision<br />

Revision<br />

51. Oversee and monitor <strong>Recovery</strong> <strong>Plan</strong><br />

implementation.<br />

511. Conduct section 7 consultation on<br />

any Federal actions that may affect<br />

<strong>Mexican</strong> spotted owls.<br />

5111. Conduct a workshop<br />

between <strong>Recovery</strong> Team and<br />

FWS consultation biologists<br />

on evaluation of projects<br />

<strong>for</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

compliance.<br />

5112. Consult programmatically<br />

on each agency’s incorporation<br />

of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

into land management<br />

policy and guidance<br />

documents.<br />

5113. Review projects <strong>for</strong> compliance<br />

with <strong>the</strong> <strong>Recovery</strong><br />

135<br />

<strong>Plan</strong>.<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

512. Form <strong>Recovery</strong> Unit Working<br />

Teams <strong>for</strong> each <strong>Recovery</strong> Unit.<br />

5121. Appoint working Team<br />

members<br />

5122. Develop charter, protocols<br />

<strong>for</strong> agreeing upon recommendations<br />

to be made to<br />

<strong>Recovery</strong> Team<br />

(e.g., voting protocols).<br />

5123. Conduct training session<br />

with <strong>Recovery</strong> Team to<br />

ensure understanding and<br />

consistent interpretation of<br />

<strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>.<br />

5124. Conduct <strong>Recovery</strong> <strong>Plan</strong><br />

implementation workshops<br />

with biologists and o<strong>the</strong>r<br />

land-management<br />

personnel.<br />

5125. Convene approximately<br />

quarterly or as needed.<br />

5126. Working Team Leaders<br />

attend all <strong>Recovery</strong> Team<br />

meetings.<br />

513. Retain <strong>Recovery</strong> Team throughout<br />

<strong>the</strong> life of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>.<br />

5131. Convene <strong>Recovery</strong> Team<br />

semi–annually <strong>for</strong> a mini<br />

mum of two years after<br />

<strong>Recovery</strong> <strong>Plan</strong> adoption.<br />

5132. Convene <strong>Recovery</strong> Team<br />

annually <strong>the</strong>reafter.<br />

52. Oversee Research<br />

521. <strong>Recovery</strong> Unit Working Teams<br />

should review and prioritize research<br />

proposals and make recommendations<br />

to <strong>the</strong> <strong>Recovery</strong> Team.<br />

522. <strong>Recovery</strong> Unit Working Teams<br />

should annually update <strong>the</strong> FWS<br />

and <strong>the</strong> <strong>Recovery</strong> Team on planned<br />

studies.


53. Review, evaluate, and revise recovery plan<br />

as appropriate.<br />

Volume I/Part IV<br />

531. <strong>Recovery</strong> Unit Working Teams<br />

should review plan implementation<br />

at least annually, reporting <strong>the</strong><br />

results to <strong>the</strong> FWS and <strong>the</strong><br />

<strong>Recovery</strong> Team.<br />

532. <strong>Recovery</strong> Unit Working Teams<br />

should review research and suggest<br />

plan revisions, if any, to <strong>the</strong> FWS<br />

and <strong>Recovery</strong> Team.<br />

54. <strong>Recovery</strong> Unit Working Teams should<br />

provide technical assistance when<br />

requested.<br />

541. Provide land managers with technical<br />

assistance in designing projects<br />

to minimize impacts on spotted<br />

owls.<br />

542. Provide technical assistance in<br />

procuring funding and logistic<br />

support <strong>for</strong> research projects.<br />

136<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

543. Provide technical assistance in<br />

developing spotted owl management<br />

plans.<br />

544. Provide technical assistance<br />

in developing conservation<br />

agreements.<br />

545. Provide o<strong>the</strong>r technical assistance<br />

as needed.<br />

55. Conduct <strong>Mexican</strong> spotted owl status<br />

reviews.<br />

56. State and private lands.<br />

561. Conduct assessment of <strong>Mexican</strong><br />

spotted owl status on State and<br />

private lands.<br />

562. Promulgate rule under 4(d) of <strong>the</strong><br />

Endangered Species Act to provide<br />

<strong>for</strong> <strong>Mexican</strong> spotted owl conserva–<br />

tion on State and private lands.


Table IV.D.1 displays estimated costs and<br />

an approximate schedule <strong>for</strong> implementing <strong>the</strong><br />

recovery tasks listed in <strong>the</strong> stepdown outline<br />

provided in IV.C. More detailed in<strong>for</strong>mation on<br />

<strong>the</strong> recommended actions is provided in Part III.<br />

The following material explains relevant details<br />

about Table IV.D.1:<br />

Task ask ask: ask This column lists specific tasks recommended<br />

in Part III. The <strong>for</strong>mat of this column is<br />

similar to that used in <strong>the</strong> stepdown outline in<br />

III.C, with each task under one of five general<br />

task categories (preceded by an Arabic numeral).<br />

In some cases “subtasks” are included if <strong>the</strong><br />

<strong>Recovery</strong> Team wished to identify specific<br />

intermediate actions to accomplish an ultimate<br />

objective. Please refer to <strong>the</strong> stepdown outline in<br />

III.C <strong>for</strong> a more detailed description of each<br />

task. Part III provides yet more detail, such as<br />

suggested methodologies and rationales.<br />

Task ask ask N NNo.<br />

N o. o.: o. This column lists <strong>the</strong> task numbers<br />

as developed in <strong>the</strong> stepdown outline (IV.C).<br />

P: This column assigns priority numbers as<br />

follows:<br />

1: 1: Tasks that must be completed to achieve<br />

<strong>the</strong> delisting criteria detailed in III.A.<br />

(Example: Population monitoring); tasks<br />

required by law (Example: Section 7<br />

consultation); and o<strong>the</strong>r tasks essential to<br />

<strong>Recovery</strong> <strong>Plan</strong> implementation (Example:<br />

amendment of agency planning documents).<br />

2: 2: Tasks that should be done to help attain<br />

<strong>the</strong> recovery objective. (Example: Restoration<br />

of degraded riparian areas).<br />

3: 3: Tasks that should be done to implement<br />

<strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> efficiently or to o<strong>the</strong>rwise<br />

enhance spotted owl management.<br />

(Example: general spotted owl inventory).<br />

Dur Dur Dur.: Dur Dur The approximate duration (in years) of<br />

each task. Items that are expected to take less<br />

Volume I/Part IV<br />

D. D. IMPLEMENTATION IMPLEMENTATION AND<br />

AND<br />

COST COST SCHEDULE<br />

SCHEDULE<br />

137<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

than one year are assigned <strong>the</strong> number “1.”<br />

Tasks that are ongoing are labeled “cont.” (continuous).<br />

Some tasks can be done to varying<br />

degrees or intensities, particularly research<br />

projects. In those cases, <strong>the</strong> duration is labeled<br />

“tbd” (to be determined).<br />

Resp. esp. PP<br />

Par PP<br />

ar arty ar ty ty: ty Assigns lead responsibility of each<br />

task to a specific party. This does not necessarily<br />

mean that <strong>the</strong> indicated entity has sole responsibility<br />

<strong>for</strong> completion of a specific task; <strong>the</strong><br />

<strong>Recovery</strong> Team recommends that agencies,<br />

Tribes, and o<strong>the</strong>rs work cooperatively on recovery<br />

tasks whenever possible.<br />

The following abbreviations are used:<br />

AA = As appropriate1 Ac = Action agency<br />

All = All involved2 AZ = State of Arizona<br />

BLM = Bureau of Land Management<br />

CO = State of Colorado<br />

DOD = Department of Defense<br />

FS = Forest Service<br />

FWS= Fish and Wildlife Service<br />

MA = Mescalero Apache<br />

MEX = Mexico<br />

NAV = Navajo<br />

NM = State of New Mexico<br />

NPS = National Park Service<br />

PI = Principle Investigator<br />

RT = <strong>Recovery</strong> Team<br />

SCA = San Carlos Apache<br />

tbd = to be determined<br />

UT = State of Utah<br />

WMA = White Mtn. Apache<br />

WT = Working Team3 1<br />

Used in situations such as under “O<strong>the</strong>r Federal<br />

agencies.”<br />

2 All parties involved in a cooperative ef<strong>for</strong>t, such as <strong>the</strong><br />

population monitoring program.<br />

3 Used both <strong>for</strong> all <strong>Recovery</strong> Unit Working Teams<br />

collectively, or <strong>for</strong> <strong>the</strong> appropriate WT <strong>for</strong> a <strong>Recovery</strong><br />

Unit.


Cost Cost Cost Estimates Estimates: Estimates The figures in this column<br />

represent <strong>the</strong> estimated costs (x$1,000) of<br />

carrying out <strong>the</strong> recommended tasks in each<br />

fiscal year (FY) indicated. Estimated costs are<br />

rounded to <strong>the</strong> nearest $1,000, i.e., a project<br />

estimated at $200 will show “0”; a project<br />

estimated at $500 will show “1”, etc. Some of<br />

<strong>the</strong> tasks assigned “NA” will be so labeled<br />

because no additional cost attributable to <strong>Mexican</strong><br />

spotted owl recovery will be incurred. These<br />

include activities that are ei<strong>the</strong>r already part of<br />

land management programs or those that can be<br />

paid <strong>for</strong> through commercial receipts (e.g. <strong>for</strong>est<br />

health enhancement/protection projects). No<br />

cost estimates are given on tasks <strong>for</strong> Mexico<br />

because <strong>the</strong> <strong>Recovery</strong> Team was unable to obtain<br />

<strong>the</strong> in<strong>for</strong>mation.<br />

Volume I/Part IV<br />

138<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Obviously, it is impossible to accurately<br />

predict <strong>the</strong> costs of many tasks. For example, <strong>the</strong><br />

cost to carry out recommended research activities<br />

can vary widely depending on <strong>the</strong> study<br />

design, <strong>the</strong> duration of <strong>the</strong> study, and o<strong>the</strong>r<br />

factors. Similarly, <strong>the</strong> fiscal year(s) under which<br />

<strong>the</strong> costs are placed may or may not be <strong>the</strong> fiscal<br />

year in which <strong>the</strong> cost is actually incurred; again,<br />

it is impossible predict when a project will be<br />

undertaken. Finally, in cases such as pre-project<br />

inventories, costs can only be estimated on a perunit<br />

basis (e.g., $1.25/acre).


139<br />

Table able IV IV.D.1 IV .D.1 Implementation and Cost Schedule


140<br />

Table able IV IV.D.1 IV .D.1 .D.1, .D.1 continued


141<br />

Table able IV IV.D.1 IV .D.1 .D.1, .D.1 continued


142<br />

Table able IV IV.D.1 IV .D.1 .D.1, .D.1 continued


143<br />

Table able IV IV.D.1 IV .D.1 .D.1, .D.1 continued


144<br />

Table able IV IV.D.1 IV .D.1 .D.1, .D.1 continued


Acronyms Acronyms and and Abbreviations<br />

Abbreviations<br />

Act - Endangered Species Act of 1973, as<br />

amended<br />

AIC - Akaike’s In<strong>for</strong>mation Criteria<br />

AK - Adaptive kernel<br />

AOU - American Ornithologists’ Union<br />

ANOVA - Analysis of variance: a statistical<br />

evaluation procedure<br />

AZ - Arizona<br />

BLM - Bureau of Land Management<br />

CJS - Cormack-Jolly-Seber: a population model<br />

CO - Colorado<br />

CSA - Coconino Study Area: a demographic<br />

study area<br />

DOD - Department of Defense<br />

EA - Environmental Assessment<br />

EIS - Environmental Impact Statement<br />

FEMAT - Forest Ecosystem Management<br />

Assessment Team<br />

FS - Forest Service (USDA Forest Service)<br />

FWS - Fish and Wildlife Service (U.S. Fish and<br />

Wildlife Service; USDI Fish and Wildlife<br />

Service)<br />

FY - Federal budget fiscal year; 1 October to 30<br />

September<br />

GIS - Geographic In<strong>for</strong>mation System<br />

GSA - Gila Study Area: a demographic study<br />

area<br />

HCA - Habitat Conservation Area<br />

ISC - Interagency Scientific Committee<br />

LMP - Land Management <strong>Plan</strong><br />

LRT - Likelihood ratio tests<br />

LSR - Late Successional Reserve<br />

MANOVA - Multivariate analysis of variance<br />

MCP - Minimum convex polygon<br />

MBTA - Migratory Bird Treaty Act<br />

NBS - National Biological Service<br />

NEPA - National Environmental Policy Act<br />

NFMA - National Forest Management Act<br />

NM - New Mexico<br />

NPS - National Park Service (USDI National<br />

Park Service)<br />

PAC - Protected Activity Center<br />

RU - <strong>Recovery</strong> Unit<br />

SISA - Sky Island Study Area<br />

SOHA - <strong>Spotted</strong> <strong>Owl</strong> Habitat Area<br />

SOMA - <strong>Spotted</strong> <strong>Owl</strong> Management Area<br />

Volume I/Glossary<br />

GLOSSARY<br />

GLOSSARY<br />

145<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

TES - Terrestrial Ecosystem Survey<br />

USDA - United States Department of Agriculture<br />

USDI - United States Department of Interior<br />

USGS - United States Geological Survey<br />

UT - Utah<br />

Ter er erms er ms<br />

Adaptive kernel (AK) - refers to a method of<br />

estimating home-range size. This method<br />

involves estimating a bivariate probability<br />

distribution from <strong>the</strong> observed animal<br />

locations, and can be used to compute<br />

<strong>the</strong> area containing a specified proportion<br />

of those locations.<br />

Adaptive management - refers to a process in<br />

which policy decisions are implemented<br />

within a framework of scientifically<br />

driven experiments to test predictions<br />

and assumptions inherent in management<br />

plans.<br />

Algorithm - a ma<strong>the</strong>matical <strong>for</strong>mula <strong>for</strong> solving<br />

a problem.<br />

Basal area - <strong>the</strong> cross-sectional area of a tree stem<br />

near its base. Generally measured at<br />

breast height (including bark).<br />

Biomass - with respect to individuals, this refers<br />

to <strong>the</strong> weight (mass) of a plant or an<br />

animal. With respect to areas or communities,<br />

refers to <strong>the</strong> total mass of living<br />

organisms in that area or community at<br />

any given time. With respect to owl diet,<br />

used to refer to <strong>the</strong> relative contribution<br />

of one species (or group) of prey animals<br />

to <strong>the</strong> overall diet.<br />

Birth-pulse population - a population assumed<br />

to have a discrete point in time during<br />

which all offspring are produced.<br />

Bonferroni confidence interval - a family of<br />

simultaneous confidence intervals in<br />

which <strong>the</strong> width of each interval is<br />

adjusted downward to account <strong>for</strong> <strong>the</strong><br />

estimation of simultaneous intervals.<br />

Basically, allows <strong>for</strong> multiple comparisons<br />

without inflating <strong>the</strong> Type I error<br />

rate.


Bosque - a discrete grove or thicket of trees,<br />

particularly in lowland or riparian areas<br />

of <strong>the</strong> Southwestern United States and<br />

Mexico; <strong>for</strong> example a cottonwood<br />

bosque or a mesquite bosque.<br />

Canopy - a layer of foliage, generally <strong>the</strong> uppermost<br />

layer, in a <strong>for</strong>est stand. Can be used<br />

to refer to mid- or understory vegetation<br />

in multi-layered stands.<br />

Canopy closure - an estimate of <strong>the</strong> amount of<br />

overhead tree cover (also canopy cover).<br />

Clearcut - an area where <strong>the</strong> entire stand of trees<br />

has been removed in one cutting.<br />

Climax species - any species that is characteristic<br />

of a plant community that through<br />

natural processes reaches <strong>the</strong> apex of its<br />

development after sufficient time. The<br />

opposite of seral species.<br />

Closed population - a population that receives<br />

no immigrants from o<strong>the</strong>r populations,<br />

and from which no individuals emigrate<br />

to o<strong>the</strong>r populations.<br />

Cohort - individuals of <strong>the</strong> same age, resulting<br />

from <strong>the</strong> same birth-pulse.<br />

Commercial <strong>for</strong>est land - <strong>for</strong>ested land deemed<br />

tentatively suitable <strong>for</strong> <strong>the</strong> production of<br />

crops of timber, that has not been<br />

withdrawn administratively from timber<br />

production (see reserved land).<br />

Confidence interval - an interval constructed<br />

around a parameter estimate, in which<br />

that estimate should occur with a specified<br />

probability, such as 95% of <strong>the</strong> time.<br />

Connectivity - an estimate of <strong>the</strong> extent to<br />

which intervening habitats connect<br />

subpopulations of spotted owls.<br />

Cordillera - a mountain range or chain.<br />

Cordilleran - of or relating to a range of mountains.<br />

Dbh - diameter at breast height, a standard<br />

measure of tree size.<br />

Demography - <strong>the</strong> quantitative analysis of<br />

population structure and trend.<br />

Demographic stochasticity - fluctuations in<br />

population size driven by random<br />

fluctuations in birth and death rates.<br />

Dispersal - The movement of organisms from<br />

<strong>the</strong>ir birth place to ano<strong>the</strong>r location<br />

where <strong>the</strong>y produce offspring.<br />

Volume I/Glossary<br />

146<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Disturbance - significant alteration of habitat<br />

structure or composition. May be natural<br />

(e.g. fire) or human-caused events (e.g.<br />

timber harvest).<br />

Early seral stage - an area that is in <strong>the</strong> early<br />

stages of ecological succession.<br />

Ecological succession - <strong>the</strong> orderly progression of<br />

an area through time from one vegetative<br />

community to ano<strong>the</strong>r in <strong>the</strong> absence of<br />

disturbance. For example, an area may<br />

proceed from grass-<strong>for</strong>b through aspen<br />

<strong>for</strong>est to mixed-conifer <strong>for</strong>est.<br />

Ecosystem - an interacting biophysical system of<br />

organisms and <strong>the</strong>ir environment.<br />

Emigration - permanent movement of individuals<br />

away from a population.<br />

Encinal - of or relating to oaks, particularly plant<br />

communities dominated by live oaks.<br />

Environmental stochasticity - random variation<br />

in environmental attributes, such as<br />

wea<strong>the</strong>r patterns or fire regimes.<br />

Even-aged <strong>for</strong>est - used to refer to <strong>for</strong>ests composed<br />

of trees with a time span of


Fuel ladder - dead or living fuels that connect<br />

fuels on <strong>the</strong> <strong>for</strong>est floor to <strong>the</strong> canopy,<br />

and promote <strong>the</strong> spread of surface fires<br />

to tree crowns.<br />

Fuel loads - <strong>the</strong> amount of combustible material<br />

present per unit area.<br />

Fuels - combustible materials.<br />

Fuelwood - wood, ei<strong>the</strong>r green or dead, harvested<br />

<strong>for</strong> purposes of cooking or space<br />

heating, and usually measured in cords.<br />

(1 cord = 128 cubic feet.)<br />

Geographical In<strong>for</strong>mation System (GIS) - a<br />

computer system capable of storing and<br />

manipulating spatial data.<br />

Gene flow - <strong>the</strong> movement of genetic material<br />

among populations.<br />

Genetic stochasticity - random changes in gene<br />

frequencies within a population, may<br />

result from factors such as inbreeding<br />

and mutation.<br />

Graminoids - any plants of <strong>the</strong> grass family in<br />

particular and also those plants in o<strong>the</strong>r<br />

families that have a grass-like <strong>for</strong>m or<br />

appearance (<strong>for</strong> example, sedges).<br />

Group-selection cutting - removal during harvest<br />

of groups of trees.<br />

Habitat - suite of existing environmental conditions<br />

required by an organism <strong>for</strong> survival<br />

and reproduction. The place where<br />

an organism typically lives.<br />

Habitat fragmentation - see fragmentation.<br />

Habitat mosaic - <strong>the</strong> mixture of habitat conditions<br />

across a landscape.<br />

Habitat type - see vegetation type.<br />

Hanging canyon - a side canyon, <strong>the</strong> mouth of<br />

which lies above <strong>the</strong> floor of a larger<br />

canyon to which it is tributary.<br />

Home range - <strong>the</strong> area used by an animal in its<br />

day-to-day activities.<br />

Immigration - <strong>the</strong> movement of individuals<br />

from o<strong>the</strong>r areas into a given area.<br />

Intermountain Region - an administrative region<br />

of <strong>the</strong> USDA Forest Service, lying<br />

between <strong>the</strong> Pacific Coastal and Rocky<br />

Mountain Ranges and including Utah,<br />

Nevada, sou<strong>the</strong>rn Idaho, and parts of<br />

Wyoming and Montana.<br />

Lambda - <strong>the</strong> finite rate of change in population<br />

size. If lambda is greater than 1, <strong>the</strong><br />

population trend is increasing; if lambda<br />

Volume I/Glossary<br />

147<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

equals 1, <strong>the</strong> population trend is stable;<br />

if lambda is less than 1, <strong>the</strong> population<br />

trend is decreasing.<br />

Land Management <strong>Plan</strong> (LMP) - a plan written<br />

<strong>for</strong> <strong>the</strong> management of a National<br />

Forest. These plans were mandated by<br />

<strong>the</strong> National Forest Management Act of<br />

1976.<br />

Late seral stage <strong>for</strong>est - a <strong>for</strong>est in <strong>the</strong> latter stages<br />

of development, usually dominated by<br />

large, old trees.<br />

Leslie matrix - a two-dimensional array of<br />

numbers representing age- or stagespecific<br />

estimates of birth and death<br />

rates, used to project population age (or<br />

stage) structure through time.<br />

Life table - ma<strong>the</strong>matical table of age- or stagespecific<br />

birth and death rates of a population.<br />

Macrohabitat - landscape-scale features that are<br />

correlated with <strong>the</strong> distribution of a<br />

species; often used to describe seral stages<br />

or discrete arrays of specific vegetation<br />

types.<br />

Madrean - pertaining to Mexico’s Sierra Madre<br />

cordillera, or to plant species or communities<br />

whose primary affinity is to that<br />

region (see also Petran).<br />

Madrean pine-oak <strong>for</strong>est - <strong>for</strong>ests in which any<br />

of several pines characterize <strong>the</strong> overstory,<br />

and midstory oaks are mostly<br />

evergreen species. Many of <strong>the</strong> dominant<br />

species are Madrean in affinity. See<br />

Marshall (1957) <strong>for</strong> descriptions. This<br />

habitat was included as Pine-oak by<br />

Fletcher and Hollis (1994).<br />

Mesic - of or relating to conditions between<br />

hydric and xeric or <strong>the</strong> specific quality of<br />

being adapted to conditions between wet<br />

and dry.<br />

Metapopulation - systems of local populations<br />

connected by dispersing individuals.<br />

Microhabitat - habitat features at a fine scale;<br />

often identifies a unique set of local<br />

habitat features.<br />

Microtine - any vole of <strong>the</strong> genus Microtus.<br />

Migration - <strong>the</strong> seasonal movement from one<br />

area to ano<strong>the</strong>r and back.<br />

Minimum convex polygon (MCP) - a method<br />

used to estimate home-range size. This


method involves <strong>for</strong>ming a polygon by<br />

connecting <strong>the</strong> outermost animal locations<br />

with a series of convex lines, <strong>the</strong>n<br />

computing <strong>the</strong> area of that polygon.<br />

Mixed-conifer <strong>for</strong>est type - overstory species in<br />

<strong>the</strong>se <strong>for</strong>ests include Rocky Mountain<br />

Douglas-fir, white fir, Rocky Mountain<br />

ponderosa pine, quaking aspen, southwestern<br />

white pine, limber pine, and<br />

blue spruce. Refer to II.C <strong>for</strong> a more<br />

precise discussion and definition of<br />

mixed-conifer <strong>for</strong>est type.<br />

Model - a representation of reality, based on a set<br />

of assumptions, that is developed and<br />

used to describe, analyze, and understand<br />

<strong>the</strong> behavior of a system of interest.<br />

Monitoring - <strong>the</strong> process of collecting in<strong>for</strong>mation<br />

to track changes of selected parameters<br />

over time.<br />

Mousing - a technique used to assess reproductive<br />

status of a pair of spotted owls.<br />

Entails feeding mice to adult owls and<br />

observing <strong>the</strong> owls’ subsequent behavior.<br />

Multi-layered (or multi-storied) stands - <strong>for</strong>est<br />

stands with >2 distinct canopy layers.<br />

Applied to <strong>for</strong>est stands that contain<br />

trees of various heights and diameters,<br />

and <strong>the</strong>re<strong>for</strong>e support foliage at various<br />

heights in <strong>the</strong> vertical profile of <strong>the</strong><br />

stand.<br />

Null hypo<strong>the</strong>sis - a hypo<strong>the</strong>sis stating that <strong>the</strong>re<br />

is no difference between units being<br />

compared.<br />

O<strong>the</strong>r <strong>for</strong>est and woodland types - vegetation<br />

types that are nei<strong>the</strong>r “restricted” or<br />

within PACs (see definitions of those<br />

terms) as to management recommendations<br />

provided in this <strong>Recovery</strong> <strong>Plan</strong>.<br />

Old growth - an old <strong>for</strong>est stand, typically<br />

dominated by large, old trees, with<br />

relatively high canopy closure and a high<br />

incidence of snags, as well as logs and<br />

o<strong>the</strong>r woody debris.<br />

Overstory - <strong>the</strong> highest limbs and foliage of a<br />

tree, and consequently extending and<br />

relating to <strong>the</strong> upper layers of a <strong>for</strong>est<br />

canopy.<br />

Pellet - a compact mass of undigested material<br />

remaining after preliminary digestion<br />

and eliminated by regurgitation ra<strong>the</strong>r<br />

Volume I/Glossary<br />

148<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

than by defecation.<br />

Peromyscid - any mouse in <strong>the</strong> genus Peromyscus<br />

of <strong>the</strong> family Muridae (<strong>for</strong>merly<br />

Cricetidae).<br />

Petran - pertaining to <strong>the</strong> Rocky Mountain area.<br />

Used to identify plant associations or<br />

species that have <strong>the</strong>ir primary affinity to<br />

<strong>the</strong> Rocky Mountain area (see also<br />

Madrean).<br />

Physiographic province - a geographic region in<br />

which climate and geology have given<br />

rise to a distinct array of land <strong>for</strong>ms and<br />

habitats.<br />

Pine-oak <strong>for</strong>est type - stands within <strong>the</strong> Pinus<br />

ponderosa and Pinus leiophylla series that<br />

exhibit a pine overstory and oak understory.<br />

Refer to II.C <strong>for</strong> <strong>the</strong>se criteria and<br />

a more precise discussion and definition<br />

of pine-oak <strong>for</strong>est type.<br />

Precommercial thinning - <strong>the</strong> practice of removing<br />

some of <strong>the</strong> smaller trees in a stand<br />

so that remaining trees will grow faster.<br />

Prescribed fire - a fire burning under specified<br />

conditions; may result from ei<strong>the</strong>r<br />

planned or unplanned ignitions.<br />

Ponderosa pine <strong>for</strong>est type - any <strong>for</strong>ested stand<br />

of <strong>the</strong> Pinus ponderosa Series not included<br />

in <strong>the</strong> pine-oak <strong>for</strong>est type<br />

definition, or any stand that qualifies as<br />

pure (i.e., any stand where a single<br />

species contributes >80 % of <strong>the</strong> basal<br />

area of dominant and codominant trees)<br />

ponderosa pine, regardless of <strong>the</strong> series or<br />

habitat (see also Eyre 1980). Refer to<br />

Part II.C <strong>for</strong> a more precise discussion<br />

and definition of ponderosa pine <strong>for</strong>est<br />

type.<br />

Population - a collection of individuals that<br />

share a common gene pool.<br />

Population density - <strong>the</strong> number of individuals<br />

per unit area.<br />

Population persistence - <strong>the</strong> capacity of a population<br />

to maintain sufficient numbers<br />

and distribution over time.<br />

Population viability - <strong>the</strong> probability that a<br />

population will persist <strong>for</strong> a specific<br />

period of time, despite demographic and<br />

environmental stochasticity.<br />

Power - with respect to statistical comparisons,<br />

refers to <strong>the</strong> probability of not making a


Type-II error.<br />

Protected Activity Center (PAC) - an area<br />

established around an owl nest (or<br />

sometimes roost) site, <strong>for</strong> <strong>the</strong> purpose of<br />

protecting that area. Management of<br />

<strong>the</strong>se areas is largely restricted to managing<br />

<strong>for</strong> <strong>for</strong>est health objectives.<br />

Protected areas - as used in this <strong>Plan</strong>, refers to<br />

areas that are protected, and where most<br />

management activities are very restricted<br />

or disallowed. Includes Protected Activity<br />

Centers.<br />

<strong>Recovery</strong> - as provided by <strong>the</strong> Endangered<br />

Species Act and its implementing regulations,<br />

<strong>the</strong> process of returning a threatened<br />

or endangered species to <strong>the</strong> point<br />

at which protection under <strong>the</strong> Endangered<br />

Species Act is no longer necessary.<br />

<strong>Recovery</strong> <strong>Plan</strong> - as provided by <strong>the</strong> Endangered<br />

Species Act, a plan <strong>for</strong> management of a<br />

threatened or endangered species that<br />

lays out <strong>the</strong> steps necessary to recover a<br />

species (see “<strong>Recovery</strong>”).<br />

<strong>Recovery</strong> Team - a team of experts appointed by<br />

<strong>the</strong> Fish and Wildlife Service whose<br />

charge is development of a <strong>Recovery</strong><br />

<strong>Plan</strong> (see “<strong>Recovery</strong> <strong>Plan</strong>”).<br />

<strong>Recovery</strong> Unit (RU) - a specific geographic area,<br />

identified mainly from physiographic<br />

provinces, used to evaluate <strong>the</strong> status of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Recruitment - <strong>the</strong> addition of individuals to a<br />

population from birth and immigration.<br />

Reserved lands - lands that have been administratively<br />

withdrawn from commercial<br />

activities, such as wilderness areas or<br />

research natural areas.<br />

Restricted Areas - as used in this <strong>Plan</strong>, refers to<br />

areas that are not protected (see Protected<br />

Areas), but where specific guidelines<br />

<strong>for</strong> management activities are<br />

proposed.<br />

Riparian - of or relating to a river; specifically<br />

applied to ecology, “riparian” describes<br />

<strong>the</strong> land immediately adjoining and<br />

directly influenced by streams. For<br />

example, riparian vegetation includes any<br />

and all plant-life growing on <strong>the</strong> land<br />

adjoining a stream and directly influenced<br />

by that stream.<br />

Volume I/Glossary<br />

149<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Riparian <strong>for</strong>ests - <strong>for</strong>ests along rivers, streams,<br />

and o<strong>the</strong>r wetland environments, typically<br />

characterized by <strong>the</strong> presence of<br />

riparian-obligate plants such as cottonwoods,<br />

willows, sycamores, or alders.<br />

Descriptions are provided by Dick-<br />

Peddie (1993) and o<strong>the</strong>rs.<br />

Rocky Mountain Region - An administrative<br />

region of <strong>the</strong> USDA Forest Service,<br />

including Colorado, Nebraska, South<br />

Dakota, and parts of Wyoming.<br />

Rotation - <strong>the</strong> planned number of years between<br />

regeneration of a <strong>for</strong>est stand and final<br />

harvest of that stand.<br />

Salvage - see sanitation salvage.<br />

Sanitation salvage - removal of dead, damaged,<br />

or susceptible trees primarily to prevent<br />

<strong>the</strong> spread of pests or pathogens and to<br />

promote <strong>for</strong>est health.<br />

Seed-tree cut - an even-aged regeneration cutting<br />

in which only a few seed trees are retained<br />

per hectare. Shelterwood cuts<br />

retain more seed trees.<br />

Seral species - any plant or animal that is typical<br />

of a seral community (stage).<br />

Seral stage - Any plant community whose plant<br />

composition is changing in a predictable<br />

way; <strong>for</strong> example, an aspen community<br />

changing to a coniferous <strong>for</strong>est community.<br />

Shelterwood cut - an even-aged regeneration<br />

cutting in which new tree seedlings are<br />

established under <strong>the</strong> partial shade of<br />

remnant seed trees.<br />

Silviculture - <strong>the</strong> practice of controlling <strong>the</strong><br />

establishment, composition, and growth<br />

of <strong>for</strong>ests.<br />

Single-tree selection cutting - a cutting method<br />

based on removal of individual trees,<br />

ra<strong>the</strong>r than groups of trees (see also<br />

group selection cutting).<br />

Sink - in a population sense, refers to a population<br />

whose death rate exceeds its birth<br />

rate. Such a population is maintained by<br />

immigration from o<strong>the</strong>r populations (see<br />

source), and is not expected to contribute<br />

to long-term population maintenance.


Slash - <strong>the</strong> residue left on <strong>the</strong> ground after<br />

logging, including logs, uprooted<br />

stumps, branches, twigs, leaves, and bark.<br />

Southwestern Region - an administrative unit of<br />

<strong>the</strong> USDA Forest Service, including<br />

Arizona and New Mexico; and an administrative<br />

unit of <strong>the</strong> USDI Fish and<br />

Wildlife Service , including Arizona,<br />

New Mexico, Texas and Oklahoma.<br />

Snag - a standing dead tree.<br />

Source - in a population sense, refers to a population<br />

where birth rate exceeds death<br />

rate. Such a population produces an<br />

excess of juveniles that can disperse to<br />

o<strong>the</strong>r populations (see sink).<br />

Spruce-fir <strong>for</strong>est type - high-elevation <strong>for</strong>ests<br />

occurring on cold sites with short growing<br />

seasons, heavy snow accumulations,<br />

and strong ecological and floristic<br />

affinities to cold <strong>for</strong>ests of higher latitudes.<br />

In general, dominant trees include<br />

Englemann spruce, subalpine and/or<br />

corkbark fir, or sometimes bristlecone<br />

pine. Refer to Part II.C <strong>for</strong> a more<br />

precise discussion and definition of<br />

spruce-fir <strong>for</strong>est type.<br />

Stand - any homogeneous area of vegetation<br />

with more or less uni<strong>for</strong>m soils, land<strong>for</strong>m,<br />

and vegetation. Typically used to<br />

refer to <strong>for</strong>ested areas.<br />

Stochastic - random or uncertain.<br />

Stringers - narrow bands of trees that extend into<br />

confined areas of suitable habitat such as<br />

in ravines.<br />

Subpopulation - a well-defined set of individuals<br />

that comprises a subset of a larger,<br />

interbreeding population (see also<br />

metapopulation).<br />

Survivorship - <strong>the</strong> proportion of newborn<br />

individuals that are alive at any given<br />

age.<br />

Team - <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

Team<br />

Technical Team - <strong>the</strong> Utah Technical Team; an<br />

interagency team charged with providing<br />

management suggestions <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl.<br />

Terrestrial Ecosystem Survey (TES) - a system of<br />

ecosystem classification, inventory,<br />

mapping, and interpretation based upon<br />

Volume I/Glossary<br />

150<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

terrestrial vegetation and environmental<br />

factors, used by <strong>the</strong> USDA Forest Service,<br />

Southwestern Region. Ecosystems<br />

are defined by combinations of potential<br />

vegetation, soils, and climates. Land is<br />

partitioned into mapping units based<br />

upon inventory data, classification, and<br />

air photo interpretation.<br />

Territory - <strong>the</strong> area that an animal defends<br />

against intruders of its own species. Not<br />

synonymous with home range, as parts<br />

of <strong>the</strong> home range are typically shared<br />

with o<strong>the</strong>r individuals.<br />

Toe clipping - a procedure by which small<br />

animals are captured alive and marked as<br />

individuals <strong>for</strong> later recapture recognition<br />

by clipping off portions of one or<br />

more toes in unique combinations.<br />

Trap-night - a standardized measurement of<br />

trapping ef<strong>for</strong>t in wildlife studies; equals<br />

one trap set <strong>for</strong> night. For example, one<br />

trap set <strong>for</strong> 10 nights and 10 traps set <strong>for</strong><br />

one night both equal 10 trap-nights.<br />

Turnover - in a population sense, refers to <strong>the</strong><br />

rate at which individuals that die are<br />

replaced by o<strong>the</strong>r individuals.<br />

Type-I error - <strong>the</strong> error made when a null<br />

hypo<strong>the</strong>sis that is true is inappropriately<br />

rejected, as when concluding that two<br />

samples from a single population come<br />

from two different populations.<br />

Type-II error - <strong>the</strong> error that is made when a null<br />

hypo<strong>the</strong>sis that is false is not rejected, as<br />

when concluding that two samples from<br />

different populations came from a single<br />

population.<br />

Understory - any vegetation whose canopy<br />

(foliage) is below, or closer to <strong>the</strong> ground<br />

than, canopies of o<strong>the</strong>r plants. The<br />

opposite of overstory.<br />

Uneven-aged management - <strong>the</strong> application of a<br />

combination of actions needed to simultaneously<br />

maintain continuous tall <strong>for</strong>est<br />

cover, recurring regeneration of desirable<br />

species, and <strong>the</strong> orderly growth and<br />

development of trees through a range of<br />

diameter or age classes. Cutting methods<br />

that develop and maintain uneven-aged<br />

stands are single-tree selection and group<br />

selection.


Vegetation types - a land classification system<br />

based upon <strong>the</strong> concept of distinct plant<br />

associations. Vegetation or habitat types<br />

(plant associations) have been documented<br />

<strong>for</strong> western <strong>for</strong>ests, and keys to<br />

<strong>the</strong>ir identification are available. The<br />

primary vegetation (or habitat) types<br />

used by <strong>Mexican</strong> spotted owls are discussed<br />

in II.C.<br />

Viability - ability of a population to persist<br />

through time (see population viability).<br />

Vital rates - collective term <strong>for</strong> age- or stage-<br />

Volume I/Glossary<br />

151<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

specific demographic rates, such as birth<br />

and death rates, of a population.<br />

Vole - any small rodent in <strong>the</strong> genus Microtus,<br />

Clethrionomys, or Phenacomys, all in <strong>the</strong><br />

family Muridae.<br />

Witches broom - a mass of profuse and densely<br />

packed twigs representing abnormal<br />

growth of a tree branch. Often results<br />

from infection by dwarf mistletoe.<br />

Xeric - of or relating to perennially dry conditions<br />

or <strong>the</strong> specific quality of being<br />

adapted to dry conditions.


Alden, P. 1969. Finding <strong>the</strong> birds in western<br />

Mexico: a guide to <strong>the</strong> states of Sonora,<br />

Sinaloa, and Nayarit. Univ. Arizona Press,<br />

Tucson. 138pp.<br />

Alexander, B. G., Jr., and F. Ronco, Jr. 1987.<br />

Classification of <strong>the</strong> <strong>for</strong>est vegetation on <strong>the</strong><br />

National Forests of Arizona and New Mexico.<br />

USDA For. Serv. Res. Note RM-469. Rocky<br />

Mtn. For. and Range Exper. Stn. Fort Collins,<br />

Colo.<br />

——., ——., E. L. Fitzhugh, and J. A. Ludwig.<br />

1984a. A classification of <strong>for</strong>est habitat types<br />

of <strong>the</strong> Lincoln National <strong>for</strong>est, New<br />

Mexico.USDA For. Serv. Gen. Tech. Rep.<br />

RM-104. Rocky Mtn. For. and Range Exper.<br />

Stn. Fort Collins, Colo.<br />

——., ——., A. S. White, and J. A. Ludwig.<br />

1984b. Douglas-fir habitat types of nor<strong>the</strong>rn<br />

Arizona. USDA For. Serv. Gen. Tech. Rep.<br />

RM-108. Rocky Mtn. For. and Range Exper.<br />

Stn. Fort Collins, Colo.<br />

American Ornithologists’ Union. 1957. Checklist<br />

of North American birds. Fifth ed. Am.<br />

Ornithologists’ Union, Washington, D.C.<br />

691pp.<br />

Ames, C. R. 1977. Wildlife conflicts in riparian<br />

management: grazing. Pages 49-51 in R. R.<br />

Johnson and D. A. Jones, eds. Importance,<br />

preservation and management of riparian<br />

habitat: a symposium. U.S. For. Serv. Gen.<br />

Tech. Rep. RM-43, Ft. Collins, Colo.<br />

Archambault, S., T. W. Swetnam, and A. M.<br />

Lynch. 1994. Western spruce budworm<br />

outbreak history in <strong>the</strong> Sacramento<br />

Mountains, New Mexico, U.S.A.<br />

Unpublished manuscript.<br />

Arizona Game and Fish Department. 1988.<br />

Threatened native wildlife in Arizona. Ariz.<br />

Game and Fish Dep., Phoenix.<br />

Attiwill, P. M. 1993. The disturbance of <strong>for</strong>est<br />

ecosystems: <strong>the</strong> ecological basis <strong>for</strong><br />

conservative management. For. Ecol. and<br />

Manage. 63:247-300.<br />

Bailey, R. G. 1980. Descriptions of <strong>the</strong><br />

ecoregions of <strong>the</strong> United States. USDA For.<br />

Serv. Misc. Pub. 1391, Intermountain Region,<br />

Ogden, Utah. 77pp.<br />

Volume I/Literature Cited<br />

LITERATURE LITERATURE CITED<br />

CITED<br />

152<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Barrowclough, G. F., and R. J. Gutiérrez. 1990.<br />

Genetic variation and differentiation in <strong>the</strong><br />

spotted owl (<strong>Strix</strong> <strong>occidentalis</strong>). Auk 107:737-<br />

744.<br />

Barrows, C. W. 1981. Roost selection by spotted<br />

owls: an adaptation to heat stress. Condor<br />

83:302-309.<br />

——. 1987. Diet shifts in breeding and<br />

nonbreeding spotted owls. J. Raptor Res.<br />

21:95-97.<br />

Bart, J., R. G. Anthony, M. Berg, J. H. Beuter,<br />

W. Elmore, J. Fay, R. J. Gutiérrez, T. H.<br />

Heintz, Jr., R. S. Holtshausen, K. Lathrop, K.<br />

Mays, R. Nafziger, M. Pagel, C. Sproul, E. E.<br />

Starkey, J. C. Tappeiner, and R. Warren. 1992.<br />

<strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted owl.<br />

Final draft. USDI Fish and Wildl. Serv.,<br />

Portland, Oregon. 194pp.<br />

Beschta, R. L., C. A. Frissell, R. Gresswell, R.<br />

Hauer, J. R. Karr, G. W. Minshall, D. A.<br />

Perry, J. J. Rhodes. 1995. Wildfire and salvage<br />

logging. Unpubl. Rep., Corvallis, Oreg.<br />

Blackburn, W. E. 1984. Impacts of grazing<br />

intensity and grazing systems on vegetation<br />

composition and production. Pages 927-984<br />

in National Research Council and National<br />

Academy of Sciences. Developing strategies<br />

<strong>for</strong> rangeland management: a report prepared<br />

by <strong>the</strong> committee on developing strategies <strong>for</strong><br />

rangeland management. Westview Press,<br />

Boulder, Colo.<br />

Block, W. M., and L. A. Brennan. 1993. The<br />

habitat concept in ornithology: <strong>the</strong>ory and<br />

application. Current Ornith. 11:35-91.<br />

——., D. M. Finch, and L. A. Brennan. 1995.<br />

Single-species versus multiple-species<br />

approaches <strong>for</strong> managment. Pages 461-476 in<br />

T. E. Martin and D. M. Finch, eds., Ecology<br />

and management of neortropical migratory<br />

birds: a syn<strong>the</strong>sis and review of critical issues.<br />

Ox<strong>for</strong>d Univ. Press, Ox<strong>for</strong>d, UK.<br />

——., K. A. With, and M. L. Morrison. 1987.<br />

On measuring bird habitat: influence of<br />

observer variability and sample size. Condor<br />

89:241-251.<br />

Bock, C. E., V. A. Saab, T. D. Rich, and D. S.<br />

Dobkin. 1993. Effects of livestock grazing on<br />

neotropical landbirds in western North


America. Pages 296-309 in D. M. Finch and<br />

P. W. Stangel, eds. Status and management of<br />

neotropical migratory birds. USDA For. Serv.<br />

Gen. Tech. Rep. RM-229, Ft. Collins, Colo.<br />

Brown, D. E., ed. 1982. Biotic communities of<br />

<strong>the</strong> American southwest-United States and<br />

Mexico. Desert <strong>Plan</strong>ts 4:1-342.<br />

——., C. H. Lowe, and C. P. Pase. 1980. A<br />

digitized systematic classification <strong>for</strong><br />

ecosystems with an illustrated summary of <strong>the</strong><br />

natural vegetation of North America. USDA<br />

For. Serv. Gen. Tech. Rep. RM-73, Ft.<br />

Collins, Colo. 93pp.<br />

Bryant, L. D. 1982. Response of livestock to<br />

riparian zone exclusion. J. Range Manage.<br />

35:780-785.<br />

Bull, E. L., R. S. Holthausen, and D. B. Marx.<br />

1990. How to determine snag density. West.<br />

J. Appl. For. 5:56-58.<br />

Burnham, K. P., and D. R. Anderson. 1992.<br />

Data-based selection of an appropriate<br />

biological model: <strong>the</strong> key to modern data<br />

analysis. Pages 16-30 in D. R. McCullough<br />

and R. H. Barrett, eds. Wildlife 2001:<br />

populations. Elsevier Applied Science, New<br />

York, N.Y.<br />

——., and W. S. Overton. 1978. Estimation of<br />

<strong>the</strong> size of a closed population when capture<br />

probabilities vary among animals. Biometrika<br />

65:625-633.<br />

——., and ——. 1979. Robust estimation of<br />

population size when capture probabilities<br />

vary among animals. Ecology 60:927-936.<br />

——., D. R. Anderson, and G. C. White. 1994.<br />

Estimation of vital rates of <strong>the</strong> nor<strong>the</strong>rn<br />

spotted owl. Colo. Coop. Fish and Wildl. Res.<br />

Unit, Colorado State Univ., Fort Collins.<br />

44pp.<br />

——., ——., ——., C. Brownie, and K. H.<br />

Pollock. 1987. Design and analysis methods<br />

<strong>for</strong> fish survival experiments based on releaserecapture.<br />

Am. Fisheries Soc. Monogr. 5:1-<br />

437.<br />

Burt, W.H. 1943. Territoriality and home range<br />

concepts as applied to mammals. J. Mammal.<br />

24:346-352.<br />

Carey, A. B. 1985. A summary of <strong>the</strong> scientific<br />

basis <strong>for</strong> spotted owl management. Pages<br />

100-114 in R. J. Gutiérrez and A. B. Carey,<br />

tech. eds. Ecology and management of <strong>the</strong><br />

Volume I/Literature Cited<br />

153<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

spotted owl in <strong>the</strong> Pacific Northwest. USDA<br />

For. Serv. Gen. Tech. Rep. PNW-185. Pacific<br />

Northwest Res. Stn. Portland, Oregon.<br />

119 pp.<br />

Chao, A. 1987. Estimating <strong>the</strong> population size<br />

<strong>for</strong> capture-recapture data with unequal<br />

catchability. Biometrics 43:783-791.<br />

——. 1988. Estimating animal abundance with<br />

capture frequency data. J. Wildl. Manage.<br />

52:295-300.<br />

——. 1989. Estimating population size <strong>for</strong><br />

sparse data in capture-recapture experiments.<br />

Biometrics 45:427-438.<br />

Choate, G. A. 1966. New Mexico’s <strong>for</strong>est<br />

resource. USDA For. Serv. Resour. Bull. INT-<br />

5. Intermountain Res. Stn. Ogden, Utah.<br />

Cirett-Galan, J. M., and Diaz, N. 1993. Estatus<br />

y distribución del búho manchado <strong>Mexican</strong>o<br />

(<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>) en Sonora, México.<br />

Centro Ecológico de Sonora, Hermosillo.<br />

66pp.<br />

Clary, W. P., and B. F. Webster. 1989. Managing<br />

grazing of riparian areas in <strong>the</strong> Intermountain<br />

Region. USDA For. Serv. Gen. Tech. Rep.<br />

INT-263, Ogden, Utah. 11pp.<br />

Colhoun, J. 1979. Predisposition by <strong>the</strong><br />

environment. Pages 75-96 in J. G. Horsfall<br />

and E. B. Cowling, eds. <strong>Plan</strong>t disease, an<br />

advanced treatise. Vol. IV: How pathogens<br />

induce disease. Academic Press,<br />

New York, N.Y.<br />

Connor, R. C., J. D. Born, A. W. Green, and R.<br />

A. O’Brien. 1990. Forest resources of Arizona.<br />

USDA For. Serv. Resour. Bull. INT-69,<br />

Ogden, Utah.<br />

Cooper, C. F. 1960. Changes in vegetation,<br />

structure, and growth of southwestern pine<br />

<strong>for</strong>ests since white settlement. Ecol. Monogr.<br />

30:129-164.<br />

Cooperrider, A. 1991. Conservation of<br />

biodiversity on western rangelands. Pages 40-<br />

53 in W. E. Hudson, ed. Landscape linkages<br />

and biodiversity. Island Press, Washington,<br />

D.C.<br />

Covington, W. W., and M. M. Moore. 1992.<br />

Postsettlement changes in natural fire regimes:<br />

implications <strong>for</strong> restoration of old-growth<br />

ponderosa pine <strong>for</strong>ests. Pages 81-99 in M. R.<br />

Kaufmann, W. H. Moir, and R. L. Bassett,<br />

eds. Old-growth <strong>for</strong>ests in <strong>the</strong> Southwest and


Rocky Mountain regions: proceedings of a<br />

workshop. U.S. For. Serv. Gen. Tech. Rep.<br />

RM-213, Ft. Collins, Colo.<br />

——., and ——. 1994a. Southwestern<br />

ponderosa <strong>for</strong>est structure: changes since<br />

Euro-American settlement. J. For.<br />

92(1):39-47.<br />

——., and ——. 1994b. Postsettlement changes<br />

in natural fire regimes and <strong>for</strong>est structure:<br />

ecological restoration of old-growth<br />

ponderosa pine <strong>for</strong>ests. J. Sustainable For.<br />

2:153-181.<br />

——., and ——. 1994c. Postsettlement changes<br />

in natural fire regimes and <strong>for</strong>est structure:<br />

ecological restoration of old-growth<br />

ponderosa pine <strong>for</strong>ests. Pages 153-181 in R.<br />

N. Sampson and D. L. Adams, eds. Assessing<br />

<strong>for</strong>est ecosytem health in <strong>the</strong> inland west.<br />

Haworth Press, Inc.<br />

Cuanalo de la Cerda, H., E. Ojeda-Trejo, A.<br />

Santos-Ocampo, and C. A. Ortíz-Solorio.<br />

1989. Provincias, regiones y subrregiones<br />

terrestres de México. Colegio de<br />

Postgraduados, Centro de Edafología,<br />

Chapingo, México.<br />

Curtis, B. F. 1960. Major geologic features of<br />

Colorado. Pages 1-30 in Geol. Soc. of Amer.,<br />

Rocky Mtn. Assoc. of Geol., Colo Sci. Soc.<br />

Daniel, T. W., J. A. Helms, and F. S. Baker.<br />

1979. Principles of silviculture. Second ed.<br />

McGraw-Hill, New York, N.Y. 500pp.<br />

Daubenmire, R. 1943. Vegetation zonation in<br />

<strong>the</strong> Rocky Mountains. Bot. Rev. 9:325-393.<br />

——. 1952. Forest vegetation of nor<strong>the</strong>rn Idaho<br />

and adjacent Washington, and its bearing on<br />

concepts of vegetation classification. Ecol.<br />

Monogr. 22:301-330.<br />

——. 1968. <strong>Plan</strong>t communities: a textbook of<br />

plant synecology. Harper and Row. New<br />

York, N.Y. 300pp.<br />

Davis, M. B. 1989. Lags in vegetation response<br />

to greenhouse warming. Climatic Change<br />

15:75-82.<br />

Dawson, W. R., J. D. Ligon, J. R. Murphy, J. P.<br />

Myers, D. Simberloff, and J. Verner. 1987.<br />

Report of <strong>the</strong> scientific advisory panel on <strong>the</strong><br />

spotted owl. Condor 89:205-229.<br />

DeBano, L. F., and L. J. Schmidt. 1989a.<br />

Improving southwestern riparian areas<br />

through watershed management. USDA For.<br />

Volume I/Literature Cited<br />

154<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Serv. Gen. Tech. Rep. RM-182. 33pp, Ft.<br />

Collins, Colo.<br />

——., and ——. 1989b. Interrelationships<br />

between watershed condition and riparian<br />

health in southwestern United States. Pages<br />

45-52 in R. E. Gresswell, B. A. Barton, and J.<br />

L. Kershner, eds. Practical approaches to<br />

riparian resource management. USDI Bur.<br />

Land Manage., Billings, Montana.<br />

DeVelice, R. L., J. A. Ludwig, W. H. Moir, and<br />

F. Ronco, Jr. 1986. A Classification of <strong>for</strong>est<br />

habitat types of Nor<strong>the</strong>rn New Mexico and<br />

Sou<strong>the</strong>rn Colorado. USDA For. Serv. Gen.<br />

Tech. Rep. RM-131. Rocky Mtn. For. and<br />

Range Exper. Stn. Fort Collins, Colo.<br />

Diaz, N., and D. Apostol. 1993. Forest<br />

landscape analysis and design: a process <strong>for</strong><br />

developing and implementing land<br />

management objectives <strong>for</strong> landscape patterns.<br />

USDA For. Serv., Pacific Northwest Region,<br />

ECO-TP-043-92. 91pp.<br />

Dickman, A. 1992. <strong>Plan</strong>t pathogens and longterm<br />

ecosystem changes. Pages 499-520 in G.<br />

C. Carroll and D. T. Wicklow, eds. The fungal<br />

community: its organization and role in <strong>the</strong><br />

ecosystem. Marcel Dekker, New York, N.Y.<br />

Dick-Peddie, W. A. 1993. New Mexico<br />

vegetation, past, present and future. Univ.<br />

New Mexico Press, Albuquerque. 244pp.<br />

Dinoor, A., and N. Eshed. 1984. The role and<br />

importance of pathogens in natural plant<br />

communities. Ann. Rev. Phytopathol.<br />

22:443-466.<br />

Dixon, G. 1991. Central Rockies Gemgyn<br />

variant of <strong>the</strong> <strong>for</strong>est vegetation simulator.<br />

Unpub. Rep. USDA For. Serv. WO-TM Serv.<br />

Center, Ft. Collins, Colo. 11pp.<br />

Duncan, R. B., and J. D. Taiz. 1992. A<br />

preliminary understanding of <strong>Mexican</strong><br />

spotted owl habitat and distribution in <strong>the</strong><br />

Chiricahua Mountains and associated sub-<br />

Mogollon mountain ranges in sou<strong>the</strong>astern<br />

Arizona. Pages 58-61 in A. M. Barton and S.<br />

A. Sloane, eds. Proc. Chiricahua Mtns. Res.<br />

Symp. Southwest Parks and Monuments<br />

Assoc., Tucson, Arizona.<br />

Dwyer, D. D., J. C. Buckhouse, and W. S. Huey.<br />

1984. Impacts of grazing intensity and<br />

specialized grazing systems on <strong>the</strong> use and<br />

value of rangeland: summary and


ecommendations. Pages 867-884 in National<br />

Research Council and National Academy of<br />

Sciences. Developing strategies <strong>for</strong> rangeland<br />

management: a report prepared by <strong>the</strong><br />

committee on developing strategies <strong>for</strong><br />

rangeland management. Westview Press,<br />

Boulder, Colo.<br />

Earhart, C. M., and N. K. Johnson. 1970. Sex,<br />

dimorphism, and food habits of North<br />

American owls. Condor 72:251-264.<br />

Eberhardt, L. L., and J. M. Thomas. 1991.<br />

Designing environmental field studies. Ecol.<br />

Monogr. 61:53-73.<br />

Edminster, C. B., H. T. Mowrer, R. L.<br />

Mathasen, T. H. Shuler, W. K. Olsen. F. G.<br />

Hawksworth. 1991. GENGYM: a variable<br />

density stand table projection system<br />

calibrated <strong>for</strong> mixed-conifer and ponderosa<br />

pine stands in <strong>the</strong> Southwest. USDA For.<br />

Serv. Res. Pap. RM-297, Rocky Mtn. For.<br />

Range Exp. Stn, Ft. Collins, Colo. 32pp.<br />

Eyre, F.H. 1980. Forest cover types of <strong>the</strong> United<br />

States and Canada. Society of American<br />

Foresters, Be<strong>the</strong>sda, Md. 148pp.<br />

Fitton, S. D. 1991. Vocal learning and call<br />

structure of male nor<strong>the</strong>rn spotted owls in<br />

northwestern Cali<strong>for</strong>nia. M.S. Thesis.<br />

Humboldt State Univ. Arcata, Calif. 32pp.<br />

Fitzhugh, E. L., W. H. Moir, J. A. Ludwig, and<br />

F. Ronco, Jr. 1987. Forest habitat types in <strong>the</strong><br />

Apache, Gila, and part of <strong>the</strong> Cibola National<br />

Forests, Arizona and New Mexico. USDA<br />

For. Serv. Gen. Tech. Rep. RM-145. Rocky<br />

Mtn. For. and Range Exper. Stn.<br />

Fort Collins, Colo.<br />

Fleischner, T. L. 1994. Ecological costs of<br />

livestock grazing in western North America.<br />

Cons. Biol. 8:629-644.<br />

Fletcher, K. W. 1990. Habitats used, abundance<br />

and distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl,<br />

<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>, on National Forest<br />

system lands. USDA For. Serv., Southwestern<br />

Region, Albuquerque, N.M. 55pp.<br />

——., and H. E. Hollis. 1994. Habitats used,<br />

abundance, and distribution of <strong>the</strong> Mex- ican<br />

spotted owl (<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>) on<br />

National Forest System lands in <strong>the</strong><br />

Southwestern Region. USDA For. Serv.,<br />

Southwestern Region, Albuquerque, N.M.<br />

Volume I/Literature Cited<br />

155<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Forest Ecosystem Management Assessment<br />

Team. 1993. Forest ecosystem management:<br />

an ecological, economic, and social<br />

assessment. Unpubl. Rep. U.S. Gov. Print.<br />

Off. (1993 794-478), Washington, D. C.<br />

Forsman, E. D. 1976. A preliminary<br />

investigation of <strong>the</strong> spotted owl in Oregon.<br />

M.S. Thesis, Oregon State Univ., Corvallis.<br />

127pp.<br />

——. 1981. Molt of <strong>the</strong> spotted owl. Auk<br />

98:735-742.<br />

——., E. C. Meslow, and H. M. Wight. 1984.<br />

Distribution and biology of <strong>the</strong> spotted owl in<br />

Oregon. Wildl. Monogr. 87:1-64.<br />

Foster, C. C., E. D. Forsman, E. C. Meslow, G.<br />

S. Miller, J. A. Reid, F. F. Wagner, A. B. Carey,<br />

and J. B. Lint. 1992. Survival and<br />

reproduction of radio-marked adult spotted<br />

owls. J. Wildl. Manage. 56:91-95.<br />

Franklin, A. B. 1992. Population regulation in<br />

nor<strong>the</strong>rn spotted owls: <strong>the</strong>oretical<br />

implications <strong>for</strong> management. Pages 815-827<br />

in D. R. McCullough and R. H. Barrett, eds.<br />

Wildlife 2001: populations. Elsevier Applied<br />

Science, New York, N.Y.<br />

——., J. P. Ward, R. J. Gutiérrez, and G. I.<br />

Gould, Jr. 1990. Density of nor<strong>the</strong>rn spotted<br />

owls in northwestern Cali<strong>for</strong>nia. J. Wildl.<br />

Manage. 54:1-10.<br />

Franklin, J. F. 1993. Preserving biodiversity:<br />

species, ecosystems, or landscapes? Ecol.<br />

Applic. 3:202-205.<br />

——. 1994. Ecosystem management: an<br />

overview. Ecosystem management:<br />

applications <strong>for</strong> sustainable <strong>for</strong>est and wildlife<br />

resources. Stevens Point, Wisc. 26pp.<br />

Ganey, J. L. 1988. Distribution and habitat<br />

ecology of <strong>Mexican</strong> spotted owls in Arizona.<br />

M.S. Thesis, Nor<strong>the</strong>rn Arizona Univ.,<br />

Flagstaff. 229 pp.<br />

——. 1990. Calling behavior of spotted owls in<br />

nor<strong>the</strong>rn Arizona. Condor 92:485-490.<br />

——., and R. P. Balda. 1989a. Distribution and<br />

habitat use of <strong>Mexican</strong> spotted owls in<br />

Arizona. Condor 91:355-361.<br />

——., and ——. 1989b. Home-range<br />

characteristics of spotted owls in nor<strong>the</strong>rn<br />

Arizona. J. Wildl. Manage. 53:1159-1165.


——., and ——. 1994. Habitat selection by<br />

<strong>Mexican</strong> spotted owls in nor<strong>the</strong>rn Arizona.<br />

Auk 111:162-169.<br />

——., and J. L. Dick, Jr. 1995. Habitat<br />

relationships of <strong>the</strong> <strong>Mexican</strong> spotted owl:<br />

Current knowledge. Chapter 2 (25pp.) in<br />

USDI 1995. Final recovery plan <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. Volume II. U.S. Fish<br />

and Wildl. Serv. Albuquerque, N.M.<br />

——., R. P. Balda, and R. M. King. 1993.<br />

Metabolic rate and evaporative water loss of<br />

<strong>Mexican</strong> spotted and great horned owls.<br />

Wilson Bull. 105:645-656.<br />

——., R. B. Duncan, and W. M. Block. 1992.<br />

Use of oak and associated woodlands by<br />

<strong>Mexican</strong> spotted owls in Arizona. Pages 125-<br />

128 in P. F. Ffolliott, G. J. Gottfried, D. A.<br />

Bennett, V. M. Hernandez C., A. Ortega-<br />

Rubio, and R. H. Hamre, eds. Ecology and<br />

management of oak and associated<br />

woodlands: perspectives in <strong>the</strong> southwestern<br />

United States and nor<strong>the</strong>rn Mexico. USDA<br />

For. Serv. Gen. Tech. Rep. RM-218, Ft.<br />

Collins, Colo.<br />

——., J. A. Johnson, R. P. Balda, and R. W.<br />

Skaggs. 1988. Status report: <strong>Mexican</strong> spotted<br />

owl. Pages 145-150 in R. L. Glinski et al., eds.<br />

Proc. Southwest Raptor Manage. Symp. and<br />

Workshop. Natl. Wildl. Feder. Sci. and Tech.<br />

Ser. 11.<br />

Grubb, T. G., and W. L. Eakle. 1988. Recording<br />

wildlife locations with <strong>the</strong> Universal<br />

Transverse Mercator (UTM) grid system.<br />

USDA For. Serv. Res. Note RM-483, Ft.<br />

Collins, Colo. 3pp.<br />

Gutiérrez, R. J. 1985. An overview of recent<br />

research on <strong>the</strong> spotted owl. Pages 39-49 in<br />

R. J. Gutiérrez and A. B. Carey, tech. eds.<br />

Ecology and management of <strong>the</strong> spotted owl<br />

in <strong>the</strong> Pacific Northwest. USDA For. Serv.<br />

Gen. Tech. Rep. PNW-185. Pacific Northwest<br />

Res. Stn. Portland, Oreg. 119 pp.<br />

——. 1989. Hematozoa from <strong>the</strong> spotted owl. J.<br />

Wildl. Dis. 25:614-618.<br />

——. 1994. Conservation planning: lessons<br />

from <strong>the</strong> spotted owl. Pages 51-58 in W. W.<br />

Covington and L. F. Debano, eds. Sustainable<br />

ecological systems: implementing an<br />

ecological approach to land management.<br />

Volume I/Literature Cited<br />

156<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

USDA For. Serv. Gen. Tech. Rep. RM-247,<br />

Ft. Collins, Colo.<br />

——, A. B. Franklin, and W. S. LaHaye. 1995.<br />

<strong>Spotted</strong> owl. in A. Poole, P. Stettenheim, and<br />

F. Gill, eds. The birds of North America.<br />

Acad. Nat. Sciences and Am. Ornithol.<br />

Union. Washington, D.C.<br />

——., ——., W. LaHaye, V. J. Meretsky, and J.<br />

P. Ward. 1985. Juvenile spotted owl dispersal<br />

in northwestern Cali<strong>for</strong>nia: preliminary<br />

results. Pages 60-65 in R. J. Gutiérrez and A.<br />

B. Carey, eds. Ecology and management of<br />

<strong>the</strong> spotted owl in <strong>the</strong> Pacific Northwest.<br />

USDA For. Serv. Gen. Tech. Rep. PNW-185,<br />

Portland, Oreg.<br />

Haack, R. A., and J. W. Byler. 1993. Insects and<br />

pathogens, regulators of <strong>for</strong>est ecosystems. J.<br />

For. 91(9):32-37.<br />

Hamer, T. E., E. D. Forsman, A. D. Fuchs, and<br />

M. L. Waters. 1992. Hybridization between<br />

barred and spotted owls. Unpubl. abstract.<br />

Raptor Res. Foundation Annu. Meeting.<br />

Hammond, E. H. 1965. Physical subdivisions.<br />

USDI Geological Survey.<br />

Hanks, J. P., and W. A. Dick-Peddie. 1974.<br />

Vegetation patterns of <strong>the</strong> White Mountains,<br />

New Mexico. Southwest. Nat. 18:371-382.<br />

——., E. L. Fitzhugh, and S. R. Hanks. 1983. A<br />

habitat type classification system <strong>for</strong><br />

ponderosa pine <strong>for</strong>ests of nor<strong>the</strong>rn Arizona.<br />

USDA For. Serv. Gen. Tech. Rep. RM-97.<br />

Rocky Mtn. For. and Range Exper. Stn. Fort<br />

Collins, Colo.<br />

Hanley, T. A., and J. L. Page. 1982. Differential<br />

effects of livestock use on habitat structure<br />

and rodent populations in Great Basin<br />

communities. Calif. Fish and Game<br />

68:160-174.<br />

Hawksworth, F. G., and D. Conklin. 1990.<br />

White pine blister rust in New Mexico. Pages<br />

43-44 in J. Hoffman and L. H. Spiegel, eds.<br />

Proc. 38th Annu. Western International For.<br />

Disease Work Conf. U.S. For. Serv., For. Pest<br />

Manage., Boise, Idaho.<br />

Hessburg, P. F., and J. S. Beatty. 1986.<br />

Incidence, severity, and growth losses<br />

associated with ponderosa pine dwarf<br />

mistletoe on <strong>the</strong> Lincoln National Forest,<br />

New Mexico. USDA For. Serv. Southwestern<br />

Region, Forest Pest Manage. R-3 86-5. 30pp.


Hobbs, R. J., and L. F. Huenneke. 1992.<br />

Disturbance, diversity, and invasion:<br />

implications <strong>for</strong> conservation. Cons. Biol.<br />

6:324-337.<br />

Hubbard, J. P. 1977. Importance of riparian<br />

ecosystems: biotic considerations. Pages 14-18<br />

in R. R. Johnson and D. A. Jones, eds.<br />

Importance, preservation and management of<br />

riparian habitat: a symposium. USDA For.<br />

Serv. Gen. Tech. Rep. RM-43, Ft. Collins,<br />

Colo.<br />

Hunter, M. L., Jr. 1991. Coping with ignorance:<br />

<strong>the</strong> coarse-filter strategy <strong>for</strong> maintaining<br />

biodiversity. Pages 266-281 in K. A. Kohm,<br />

ed. Balancing on <strong>the</strong> brink of extinction.<br />

Island Press, Washington, D.C.<br />

Hunter, J. E., R. J. Gutiérrez, A. B. Franklin,<br />

and D. Olson. 1994. Ectoparasites of <strong>the</strong><br />

spotted owl. J. Raptor Res. 28:232-235.<br />

Hurlburt, S. H. 1984. Pseudoreplication and <strong>the</strong><br />

design of ecological field experiments. Ecol.<br />

Monogr. 54:187-211.<br />

Jensen, M. E., and P. S. Bourgeron. 1993.<br />

Ecosystem management: principles and<br />

applications. USDA For. Serv., Nor<strong>the</strong>rn<br />

Region. 388pp.<br />

Johnsgard, P. A. 1988. North American owls:<br />

biology and natural history. Smithsonian Inst.<br />

Press, Washington, D.C. 295pp.<br />

Johnson, J. A., and T. H. Johnson. 1985. The<br />

status of <strong>the</strong> spotted owl in nor<strong>the</strong>rn New<br />

Mexico. Unpubl. rep. New Mexico Dep.<br />

Game and Fish, Santa Fe. 39pp.<br />

Johnson, M. 1994. Changes in Southwestern<br />

<strong>for</strong>ests: stewardship implications. J. For.<br />

92(12):16-19.<br />

Kauffman, J. B., and W. C. Krueger. 1984.<br />

Livestock impacts on riparian ecosystems and<br />

streamside management implications, a<br />

review. J. Range Manage. 37:430-438.<br />

——., ——., and M. Vavra. 1983. Effects of late<br />

season cattle grazing on riparian plant<br />

communities. J. Range Manage. 36:685-691.<br />

Kaufmann, M. R., R. T. Graham, D. A. Boyce,<br />

W. H. Moir, L. Perry, R. T. Reynolds, R. L.<br />

Bassett, P. Mehlhop, C. B. Edminster, W. M.<br />

Block, and P. S. Corn. 1994. An ecological<br />

basis <strong>for</strong> ecosystem management. USDA For.<br />

Serv. Gen. Tech. Rep. RM-246, Ft. Collins,<br />

Colo. 22pp.<br />

Volume I/Literature Cited<br />

157<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Keitt, T. H., A. B. Franklin, and D. L. Urban.<br />

1995. Landscape analysis and metapopulation<br />

structure. Chapter 3 (16pp.) in USDI 1995.<br />

<strong>Mexican</strong> spotted owl recovery plan.<br />

Volume II. USDI Fish and Wildl. Serv.<br />

Albuquerque, N.M.<br />

Kendall, W. L., and K. H. Pollock. 1992. The<br />

robust design in capture-recapture studies.<br />

Pages 31-43 in D. R. McCullough and R. H.<br />

Barrett, eds. Willdife 2001: populations.<br />

Elsevier Applied Sci., New York, N.Y.<br />

Kennedy, C. E. 1977. Wildlife conflicts in<br />

riparian management: water. Pages 52-58 in<br />

R. R. Johnson and D. A. Jones, eds.<br />

Importance, preservation and management of<br />

riparian habitat: a symposium. USDA For.<br />

Serv. Gen. Tech. Rep. RM-43,<br />

Ft. Collins, Colo.<br />

Kertell, K. 1977. The spotted owl at Zion<br />

National Park, Utah. West. Birds 8:147-150.<br />

Knauer, K. H. 1988. Forest health through<br />

silviculture and integrated pest management, a<br />

strategic plan. USDA For. Serv. 26pp.<br />

Knopf, F. L., R. R. Johnson, T. Rich, F. B.<br />

Samson, and R. C. Szaro. 1988 Conservation<br />

of riparian ecosystems in <strong>the</strong> United States.<br />

Wilson Bull. 100:272-284.<br />

Kristan, D. M., R. J. Gutiérrez, A. B. Franklin,<br />

and W. S. LaHaye. In prep. Comparative<br />

morphology of <strong>the</strong> spotted owl: subspecific<br />

and regional variation. Auk.<br />

Kroel, K. and P. J. Zwank. 1991. Home range<br />

and habitat use characteristics of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl in <strong>the</strong> sou<strong>the</strong>rn Sacramento<br />

Mountains, New Mexico. Unpubl. rep.<br />

submitted to USDA For. Serv., Lincoln<br />

National Forest. New Mexico Coop. Fish and<br />

Wildl. Unit, Las Cruces.<br />

Küchler, A. W. 1964. The potential natural<br />

vegetation of <strong>the</strong> conterminous United States.<br />

Amer. Geogr. Soc. Spec. Publ. 361. 116pp.<br />

Larson, M., and W. H. Moir. 1986. Forest and<br />

woodland habitat types (plant associations) of<br />

sou<strong>the</strong>rn New Mexico and central Arizona<br />

(north of <strong>the</strong> Mogollon Rim). USDA For.<br />

Serv. Southwestern Region. Albuquerque,<br />

N.M.<br />

Laymon, S. A. 1991. Diurnal <strong>for</strong>aging by<br />

spotted owls. Wilson Bull. 103:138-140.


Layser, E. F., and G. H. Schubert. 1979.<br />

Preliminary classification <strong>for</strong> <strong>the</strong> coniferous<br />

<strong>for</strong>est and woodland series of Arizona and<br />

New Mexico. USDA For. Serv. Res. Pap. RM-<br />

208. Rocky Mtn. For. and Range Exper. Stn.<br />

Fort Collins, Colo.<br />

Lebreton, J. D., K. P. Burnham, J. Clobert, and<br />

D. R. Anderson. 1992. Modeling survival and<br />

testing biological hypo<strong>the</strong>ses using marked<br />

animals: a unified approach with case studies.<br />

Ecol. Monogr. 62:67-118.<br />

Ledig, F. T. 1992. Human impacts on genetic<br />

diversity in <strong>for</strong>est ecosystems. Oikos 63:87-<br />

108.<br />

Lefebvre, L. W., D. L. Otis, and N. R. Holler.<br />

1982. Comparison of open and closed models<br />

<strong>for</strong> cotton rat population estimates. J. Wildl.<br />

Manage. 46:156-163.<br />

Marsden, M. A., C. G. I. Shaw, and M.<br />

Morrison. 1993. Simulation of management<br />

options <strong>for</strong> stands of Southwestern ponderosa<br />

pine attacked by Armillaria root disease and<br />

dwarf mistletoe. USDA For. Serv. Res. Pap.<br />

RM-308, Ft. Collins, Colo. 5pp.<br />

Maser, C., J. M. Trappe, and D. C. Ure. 1978.<br />

Implications of small mammal myco- phagy<br />

to <strong>the</strong> management of western coniferous<br />

<strong>for</strong>ests. Trans. North Am. Wildl. Nat. Resour.<br />

Conf. 43:78-88.<br />

May, C. A., M. Z. Peery, R. J. Gutiérrez, M. E.<br />

Seamans, and D.R. Olson. In press.<br />

Feasability of a random quardat study design<br />

to estimate changes in density of <strong>Mexican</strong><br />

spotted owls. U. S. For. Serv. Res. Pap. Ft.<br />

Collins, Colo.<br />

McDonald, C. B., J. Anderson, J. C. Lewis, R.<br />

Mesta, A. Ratzlaff, T. J. Tibbitts, and S. O.<br />

Williams. 1991. <strong>Mexican</strong> spotted owl (<strong>Strix</strong><br />

<strong>occidentalis</strong> <strong>lucida</strong>) status report. USDI Fish<br />

and Wildl. Serv., Albuquerque, N.M.<br />

McLaughlin, S. P. 1986. Floristic analysis of <strong>the</strong><br />

southwestern United States. Great Basin Nat.<br />

46:46-65.<br />

Medin, D. E., and W. P. Clary. 1990. Bird and<br />

small mammal populations in a grazed and<br />

ungrazed riparian habitat in Idaho. USDA<br />

For. Serv. Res. Pap. INT-425, Ogden,<br />

Utah. 8pp.<br />

Medina, A. L. 1986. Riparian plant<br />

communities of <strong>the</strong> Fort Bayard watershed in<br />

Volume I/Literature Cited<br />

158<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

southwestern New Mexico. Southwestern Nat.<br />

31:345-359.<br />

Meslow, E. C. 1993. <strong>Spotted</strong> owl protection:<br />

unintentional evolution toward ecosystem<br />

management. Endang. Species Update<br />

10:34-38.<br />

Milchunas, D. G., and W. K. Lauenroth. 1993.<br />

Quantitative effects of grazing on vegetation<br />

and soils over a global range of environments.<br />

Ecol. Monogr. 63:327-366.<br />

Miller, G. S. 1989. Dispersal of juvenile<br />

nor<strong>the</strong>rn spotted owls in western Oregon.<br />

M.S. Thesis, Oregon State Univ., Corvallis.<br />

Minckley, W. L., and T. O. Clark. 1984.<br />

Formation and destruction of a Gila River<br />

mesquite bosque community. Desert <strong>Plan</strong>ts<br />

6:23-30.<br />

——., and J. N. Rinne. 1985. Large woody<br />

debris in hot-desert streams: an historical<br />

review. Desert <strong>Plan</strong>ts 7:142- 153.<br />

Moen, C. A., A. B. Franklin, and R. J. Gutiérrez.<br />

1991. Age determination of subadult<br />

nor<strong>the</strong>rn spotted owls in northwest<br />

Cali<strong>for</strong>nia. Wildl. Soc. Bull. 19:489-493.<br />

Moir, W. H. 1993. Alpine tundra and coniferous<br />

<strong>for</strong>est. Pages 47-84 in W. A. Dick-Peddie.<br />

New Mexico vegetation, past, present and<br />

future. Univ. of New Mexico Press,<br />

Albuquerque. 244pp.<br />

——., and J. A. Ludwig. 1979. A classification<br />

of spruce-fir and mixed-conifer habitat types<br />

of Arizona and New Mexico. USDA For. Serv.<br />

Res. Pap. RM-207. Rocky Mtn. For. and<br />

Range Exper. Stn. Fort Collins, Colo.<br />

Monson, G., and A. R. Phillips. 1981.<br />

Annotated checklist of <strong>the</strong> birds of Arizona.<br />

Univ. of Arizona Press, Tucson. 240pp.<br />

Moody, R., L. Buchanan, R. Melcher, and H.<br />

Wistrand. 1992. Fire and <strong>for</strong>est health. USDA<br />

For. Serv. Southwestern Region. Albuquerque,<br />

N.M. 23pp.<br />

Murphy, D. D., and B. R. Noon. 1991. Coping<br />

with uncertainty in wildlife biology. J. Wildl.<br />

Manage. 55:773-782.<br />

Nelson, E. W. 1903. Descriptions of new birds<br />

from sou<strong>the</strong>rn Mexico. Proc. Biol. Soc. Wash.<br />

16:151-160.<br />

Nice, M. M. 1941. The role of territory in bird<br />

life. Am. Midl. Nat. 26:441-487.


Nichols, J. D. 1991. Science, population<br />

ecology, and <strong>the</strong> management of <strong>the</strong> American<br />

black duck. J. Wildl. Manage. 55:790-799.<br />

Noon, B. R., L. L. Eberhardt, A. B. Franklin, J.<br />

D. Nichols, and K. H. Pollock. 1993. A<br />

survey design to estimate <strong>the</strong> number of<br />

territorial spotted owls in Olympic National<br />

Park. Final Rep. submitted to Olympic Natl.<br />

Park. 31pp.<br />

——., K. S. McKelvey, D. W. Lutz, W. S.<br />

LaHaye, R. J. Gutiérrez, and C. A. Moen.<br />

1992. Estimates of demographic parameters<br />

and rates of population change. Pages 175-<br />

186 in J. Verner, K. S. McKelvey, B. R. Noon,<br />

R. J. Gutiérrez, G. I. Gould, Jr., and T. W.<br />

Beck, eds. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status.<br />

USDA For. Serv. Gen. Tech. Rep. PSW-133,<br />

Albany, Calif.<br />

——., and ——. 1992. Stability properties of<br />

<strong>the</strong> spotted owl metapopulation in sou<strong>the</strong>rn<br />

Cali<strong>for</strong>nia. Pages 187-206 in J. Verner, K. S.<br />

McKelvey, B. R. Noon, R. J. Gutiérrez, G. I.<br />

Gould, Jr., and T. W. Beck, eds. The<br />

Cali<strong>for</strong>nia spotted owl: a technical assessment<br />

of its current status. USDA For. Serv. Gen.<br />

Tech. Rep. PSW-133, Albany, Calif.<br />

Otis, D. L., K. P. Burnham, G. C. White, and<br />

D. R. Anderson. 1978. Statistical inference<br />

from capture data on closed animal<br />

populations. Wildl. Monogr. 62:1-135.<br />

Pase, C. P., and E. F. Layser. 1977. Classification<br />

of riparian habitat in <strong>the</strong> southwest. Page 5-9<br />

in R. R. Johnson and D. A. Jones, tech.<br />

coords. Importance, preservation, and<br />

management of riparian habitat: A<br />

symposium. USDA For. Serv. Gen. Tech. Rep.<br />

RM-43. Rocky Mtn. For. and Range Exper.<br />

Stn. Fort Collins, Colo.<br />

Paton, P. W. C., C. J. Zabel, D. L. Neal, G. N.<br />

Steger, N. G. Tilghman, and B. R. Noon.<br />

1991. Effects of radio tags on spotted owls. J.<br />

Wildl. Manage. 55:617-622.<br />

Pearson, G. A. 1931. Forest types in <strong>the</strong><br />

southwest as determined by climate and soil.<br />

USDA For. Serv. Tech. Bull. 247.<br />

Washington, D.C.<br />

Pfister, R. D. 1989. Basic concepts of using<br />

vegetation to build a site classification system.<br />

Pages 22-28 in D. E. Ferguson, P. Morgan,<br />

Volume I/Literature Cited<br />

159<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

and F. D. Johnson, compilers. Proceedingsland<br />

classifications based on vegetation:<br />

Applications <strong>for</strong> resource management.<br />

USDA For. Serv. Gen. Tech. Rep. INT-257.<br />

Intermountain Res. Stn., Ogden, Utah.<br />

Platt, J. R. 1964. Strong inference. Science<br />

46:347-353.<br />

Platts, W.S. 1990.Managing fisheries and<br />

wildlife on rangelands grazed by livestock: a<br />

guidance and reference document <strong>for</strong><br />

biologists. Nevada Dep. of Wildl. 448pp.<br />

Plummer, F. G., and M. G. Gowsell. 1904.<br />

Forest conditions in <strong>the</strong> Lincoln Forest,<br />

Reserve, NM. USDI, Geol. Surv. Prof.<br />

Pap. 33.<br />

Pollock, K. H. 1982. A capture-recapture design<br />

robust to unequal probability of capture. J.<br />

Wildl. Manage. 37:757-760.<br />

——., J. D. Nichols, C. Brownie, and J. E.<br />

Hines. 1990. Statistical inference <strong>for</strong> capturerecapture<br />

experiments. Wildl. Monogr.<br />

107:1-97.<br />

Popper, K. R. 1965. The logic of scientific<br />

discovery. Harper and Row, New York, N.Y.<br />

Powell, D. S., J. L. Faulkner, D. Darr, Z. Zhu,<br />

and D. W. MacCleery. 1993. Forest resources<br />

of <strong>the</strong> United States. USDA For. Serv., Rocky<br />

Mtn. For. Rng. Exp. Stn., Ft. Collins, Colo.<br />

Reynolds, R. T. 1990. Distribution and habitat<br />

of <strong>Mexican</strong> spotted owls in Colorado:<br />

preliminary results. USDA For. Serv., Rocky<br />

Mtn. For. Rng. Exp. Stn., Laramie, Wyoming.<br />

10pp.<br />

——. 1993. Ecology and dispersal of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl in Colorado, FY 1993-<br />

94. Research Proposal. USDA For. Serv.,<br />

Rocky Mtn. For. Rng. Exp. Stn., Fort Collins,<br />

Colo.<br />

——., R. T. Graham, M. H. Reiser, R. L.<br />

Bassett, P. L. Kennedy, D. A. Boyce, Jr., G.<br />

Goodwin, R. Smith, and E. L. Fisher. 1992.<br />

Management recommendations <strong>for</strong> <strong>the</strong><br />

nor<strong>the</strong>rn goshawk in <strong>the</strong> southwestern United<br />

States. USDA For. Serv. Gen. Tech. Rep. RM-<br />

217. Ft. Collins, Colo. 90pp.<br />

Rickard, W. H., and C. E. Cushing. 1982.<br />

<strong>Recovery</strong> of streamside woody vegetation after<br />

exclusion of livestock grazing. J. Range<br />

Manage. 35:360-361.


Ridgway, R. 1914. The birds of north and<br />

middle America. U.S. Natl. Mus. Bull. 50:1-<br />

882.<br />

Rinkevich, S. E. 1991. Distribution and habitat<br />

characteristics of <strong>Mexican</strong> spotted owls in<br />

Zion National Park, Utah. M.S. Thesis,<br />

Humboldt State Univ., Arcata, Calif. 62pp.<br />

Romesburg, H. C. 1981. Wildlife science:<br />

gaining reliable knowledge. J. Wildl. Manage.<br />

45:527-534.<br />

——. 1991. On improving <strong>the</strong> natural resources<br />

and environmental sciences. J. Wildl. Manage.<br />

55:744-756.<br />

Schulz, T. T., and W. C. Leininger. 1990.<br />

Differences in riparian vegetation structure<br />

between grazed areas and exclosures. J. Range<br />

Manage. 43:295-299.<br />

——., and ——. 1991. Nongame wildlife<br />

communities in grazed and ungrazed montane<br />

riparian sites. Great Basin Nat. 51:286-292.<br />

Seamans, M. E., and R. J. Gutiérrez. In press.<br />

Breeding habitat ecology of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl in <strong>the</strong> Tularosa Mountains, New<br />

Mexico. Condor.<br />

Short, L. L. 1965. Hybridization in <strong>the</strong> flickers<br />

(Colaptes) of North America. Bull. Amer.<br />

Mus. Nat. Hist. 129:309-428.<br />

——. 1972. Hybridization, taxonomy and avian<br />

evolution. Ann. Missouri Bot. Gard.<br />

59:447-453.<br />

Skaggs, R. W., and R. J. Raitt. 1988. A spotted<br />

owl inventory of <strong>the</strong> Lincoln National Forest,<br />

Sacramento Division, 1988. Unpubl. rep.<br />

New Mexico Dep. Game and Fish, Santa Fe.<br />

12pp.<br />

Skovlin, J. M. 1984. Grazing impacts on<br />

wetlands and riparian habitat: a review of our<br />

knowledge. Pages 1001-1004 in National<br />

Research Council and National Academy of<br />

Sciences. Developing strategies <strong>for</strong> rangeland<br />

management: a report prepared by <strong>the</strong><br />

committee on developing strategies <strong>for</strong><br />

rangeland management. Westview Press,<br />

Boulder, Colo.<br />

Smith, D. M. 1986. The practice of silviculture.<br />

Second ed. John Wiley and Sons, New York,<br />

N.Y. 527pp.<br />

Sovern, S. G., E. D. Forsman, B. L. Biswell, D.<br />

N. Rolph, and M. Taylor. 1994. Diurnal<br />

Volume I/Literature Cited<br />

160<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

behavior of <strong>the</strong> spotted owl in Washington.<br />

Condor 96:200-202.<br />

Spencer, J. S., Jr. 1966. Arizona’s <strong>for</strong>ests. USDA<br />

For. Serv. Resour. Bull. INT-6, Ogden, Utah.<br />

Stevens, R. E., and H. W. Flake, Jr. 1974. A<br />

roundheaded pine beetle outbreak in New<br />

Mexico. USDA For. Serv. Res. Note RM-259,<br />

Ft. Collins, Colo. 4pp.<br />

SWCA, Inc. 1992. Nest tree and nest tree<br />

habitat characteristics of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl in Arizona and New Mexico. Unpubl. rep.<br />

USDA For. Serv., Southwestern Region,<br />

Albuquerque, N.M.<br />

Swetnam, T. W., and J. L. Bettancourt. 1990.<br />

Fire--sou<strong>the</strong>rn oscillation relations in<br />

southwestern United States. Science<br />

249:1017-1020.<br />

——., and A. M. Lynch. 1993. Multicentury,<br />

regional-scale patterns of western spruce<br />

budworm outbreaks. Ecol. Monogr. 63:399-<br />

424.<br />

Szaro, R. C. 1989. Riparian <strong>for</strong>est and scrubland<br />

community types of Arizona and New<br />

Mexico. Desert <strong>Plan</strong>ts 9: 70-138.<br />

Tarango, L. A., R. Valdez, and P. J. Zwank.<br />

1994. <strong>Mexican</strong> spotted owl distribution and<br />

habitat characterizations in southwestern<br />

Chihuahua, Mexico. Final report, Contract<br />

No. 80-516.6-66. New Mexico Dep. Game<br />

and Fish, Santa Fe. 69pp.<br />

Thomas, C. D. 1990. What do real population<br />

dynamics tell us about minimum viable<br />

population sizes. Conserv. Biol. 4:324-327.<br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner. 1990. A<br />

conservation strategy <strong>for</strong> <strong>the</strong> spotted owl. U.S.<br />

Gov. Print. Off., Washington, D.C. 427 pp.<br />

——., M. G. Raphael, R. G. Anthony, E. D.<br />

Forsman, A. G. Gunderson, R. S.<br />

Holthausen, B. G. Marcot, G. H. Reeves, J.<br />

R. Sedell, and D. M. Solis. 1993. Viability<br />

assessments and management considerations<br />

<strong>for</strong> species associated with late- successional<br />

and old-growth <strong>for</strong>ests of <strong>the</strong> Pacific<br />

Northwest. USDA For. Serv., Portland, Oreg.<br />

529pp.<br />

Thompson, S. K. 1992. Sampling. John Wiley<br />

and Sons, New York, N.Y. 343pp.<br />

Thornbury, W. 1965. Regional geomorphology<br />

of <strong>the</strong> United States. John Wiley and Sons,


New York. N.Y. 609pp.<br />

Thrailkill, J., and M. A. Bias. 1989. Diets of<br />

breeding and nonbreeding Cali<strong>for</strong>nia spotted<br />

owls. J. Raptor Res. 23:39-41.<br />

USDA Forest Service. 1983. Silvicultural systems<br />

<strong>for</strong> <strong>the</strong> major <strong>for</strong>est types of <strong>the</strong> United States.<br />

Agri. Handbook 445. U. S. Govt. Printing<br />

Office. Washington, D.C. vii, 114pp.<br />

——. 1990. Management guidelines and<br />

inventory and monitoring protocols <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl in <strong>the</strong> Southwestern<br />

Region. Federal Register 55:27278-27287.<br />

——. 1991. General ecosystem survey, general<br />

terrestrial vegetation, existing vegetation.<br />

USDA For. Serv. Southwestern Region and<br />

Technology Application Center, Univ. New<br />

Mexico, Albuquerque. Map at scale<br />

1:1,000,000.<br />

——. 1992. Recommended old growth<br />

definitions and descriptions and old growth<br />

allocation procedure. Southwestern Region.<br />

Albuquerque, N.M.<br />

——. 1993a. Facts about <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl. USDA For. Serv., Southwestern Region,<br />

Albuquerque, N.M.<br />

——. 1993b. Changing conditions in<br />

southwestern <strong>for</strong>ests and implications on land<br />

stewardship. USDA For. Serv., Southwestern<br />

Region, Albuquerque, N.M.<br />

——. 1993c. Forest health restoration initiative.<br />

USDA For. Serv., Southwestern Region,<br />

Albuquerque, N.M. 19pp.<br />

——. 1994. Biological assessment and<br />

evaluation: livestock grazing. Unpubl. Rep.,<br />

USDA For. Serv., Southwest Regional Office,<br />

Albuquerque, N.M. 12pp.<br />

USDI. 1994. The impact of federal programs on<br />

wetlands, Vol.II, A report to Congress by <strong>the</strong><br />

Secretary of <strong>the</strong> Interior. Washington, D.C.<br />

USDI Fish and Wildlife Service. 1992. <strong>Recovery</strong><br />

plan <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted owl. Final draft.<br />

U.S. Fish and Wildl. Serv., Portland, Oreg.<br />

194pp.<br />

USDI Fish & Wildlife Service. 1993.<br />

Endangered and threatened wildlife and<br />

plants; final rule to list <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl as a threatened species. Federal Register<br />

58: 14248-14271.<br />

USDI Geological Survey. 1970. The national<br />

atlas of <strong>the</strong> United States of America.<br />

Volume I/Literature Cited<br />

161<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Washington, D.C. 417pp.<br />

Utah Division of Wildlife Resources. 1987.<br />

Native Utah wildlife species of special<br />

concern. Utah Div. Wildl. Resour., Salt Lake<br />

City.<br />

——. 1992. Native Utah wildlife species of<br />

special concern, draft. Utah Div. Wildl.<br />

Resour., Salt Lake City.<br />

Utah <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> Technical Team.<br />

1994. Suggestions <strong>for</strong> management of<br />

<strong>Mexican</strong> spotted owls in Utah. Utah Div.<br />

Wildl. Resour., Salt Lake City. 103pp.<br />

Van Hooser, D. D., D. C. Collins, and R. A.<br />

O’Brien. 1992. Forest resources of New<br />

Mexico. USDA For. Serv. Inter. Res. Stn.,<br />

Ogden, Utah. 80pp.<br />

Verner, J., K. S. McKelvey, B. R. Noon, R. J.<br />

Gutiérrez, G. I. Gould, Jr., and T. W. Beck,<br />

eds. 1992a. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status.<br />

USDA For. Serv. Gen. Tech. Rep. PSW-133,<br />

Albany, Calif. 285pp.<br />

——., ——., ——., ——., ——., and ——.<br />

1992b. Assessment of <strong>the</strong> current status of <strong>the</strong><br />

Cali<strong>for</strong>nia spotted owl, with<br />

recommendations <strong>for</strong> management. Pages 3-<br />

26 in J. Verner, K. S. McKelvey, B. R. Noon,<br />

R. J. Gutierrez, G. I. Gould, Jr., and T. W.<br />

Beck, eds. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status.<br />

USDA For. Serv. Gen. Tech. Rep. PSW-133,<br />

Albany, Calif.<br />

Walker, L. W. 1974. The book of owls. Alfred A.<br />

Knopf, New York, N.Y. 255pp.<br />

Ward, J. P., Jr., and W. M. Block.1995. <strong>Mexican</strong><br />

spotted owl prey ecology. Chapter 5 (48pp.)<br />

in USDI Fish and Wildlife Service. 1995.<br />

<strong>Mexican</strong> spotted owl recovery plan.<br />

Volume II. Albuquerque, N.M.<br />

——., A. B. Franklin, S. Rinkevich, and F.<br />

Clemente. 1995. Distribution and abundance.<br />

Chapter 1 (14pp.) in USDI Fish and Wildlife<br />

Service. 1995. <strong>Mexican</strong> spotted owl recovery<br />

plan. Volume II. Albuquerque, N.M.<br />

——., ——., and R. J. Gutiérrez. 1991. Using<br />

search time and regression to estimate<br />

abundance of territorial spotted owls. Ecol.<br />

Applic. 1:207-214.<br />

West, N. E. 1983. Western intermountain<br />

sagebrush steppe. Pages 351-374 in N. E.


West, ed. Temperate deserts and semi-deserts.<br />

Elsevier Press, New York, N.Y.<br />

White, G. C., D. R. Anderson, K. P. Burnham,<br />

and D. L. Otis. 1982. Capture-recapture and<br />

removal methods <strong>for</strong> sampling closed<br />

populations. USDE Los Alamos National<br />

Lab., Los Alamos, N.M. 235pp.<br />

White, G. C, A. B. Franklin, and J. P. Ward, Jr.<br />

1995. Population biology. Chapter 2 (25pp.)<br />

in USDI Fish and Wildlife Service. 1995.<br />

<strong>Mexican</strong> spotted owl recovery plan.<br />

Volume II. Albuquerque, N.M.<br />

Wykoff, W. R., N. L. Crookston, and A. R.<br />

Stage. 1982. User’s guide to stand prognosis<br />

model. USDA For. Serv. Gen. Tech. Rep.<br />

INT-133, Intermounain Res. Stn., Ogden,<br />

Utah. 112pp.<br />

Willey, D. W. 1993. Home-range characteristics<br />

and juvenile dispersal ecology of <strong>Mexican</strong><br />

spotted owls in sou<strong>the</strong>rn Utah. Unpubl. rep.<br />

Utah Div. Wildl. Resour., Salt Lake City.<br />

Williams, J. L. 1986. New Mexico in maps.<br />

Univ. New Mexico Press, Albuquerque.<br />

409pp.<br />

Williams, J. T. 1994. Fire’s role in support of<br />

ecosystem management. The Biswell<br />

Symposium, Walnut Creek, Calif. USDA For.<br />

Serv., Fire and Aviation Management. 4pp.<br />

Williams, S. O., III, and R. W. Skaggs. 1993.<br />

Distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl (<strong>Strix</strong><br />

<strong>occidentalis</strong> <strong>lucida</strong>) in Mexico. Unpubl. rep.,<br />

New Mexico Dep. Game and Fish, Sante Fe.<br />

38pp.<br />

Volume I/Literature Cited<br />

162<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Wilson, E. D. 1962. A resume of <strong>the</strong> geology of<br />

Arizona. Univ. Arizona Press, Tucson. 140pp.<br />

Young, K. E., A. B. Franklin, and J. P. Ward, Jr.<br />

1993. Infestation of nor<strong>the</strong>rn spotted owls by<br />

hippoboscid (Diptera) flies in nortwestern<br />

Cali<strong>for</strong>nia. J. Wildl. Dis. 29:278-283.<br />

Youngblood, A. P., and R. L. Mauk. 1985.<br />

Coniferous <strong>for</strong>est habitat types of central and<br />

sou<strong>the</strong>rn Utah. USDA For. Serv. Gen. Tech.<br />

Rep. INT-187. Intermountain Res. Stn.<br />

Ogden, Utah.<br />

Zimmerman, G. T., and L. F. Neuenschwander.<br />

1984. Livestock grazing influences on<br />

community structure, fire intensity, and fire<br />

frequency within <strong>the</strong> Douglas-fir/ninebark<br />

habitat type. J. Range Manage. 37:104-110.<br />

Zippin, C. 1956. An evaluation of <strong>the</strong> removal<br />

method of estimating animal populations.<br />

Biometrics 12:163-189.<br />

——. 1958. The removal method of population<br />

estimation. J. Wildl. Manage. 22:82-90.<br />

Zwank, P. J., K. W. Kroel, D. M. Levin, G. M.<br />

Southward, and R. C. Rommé. 1994. Habitat<br />

characteristics of <strong>Mexican</strong> spotted owls in<br />

southwestern New Mexico. J. Field Ornithol.<br />

65: 324-334.


THE THE RECOVERY RECOVERY TEAM<br />

TEAM<br />

AND AND ASSOCIATES<br />

ASSOCIATES<br />

RECOVERY RECOVERY TEAM:<br />

TEAM:<br />

William M. Block, Team Leader;<br />

Wildlife Biologist.<br />

Education: B.A., Economics, San Diego State<br />

University, 1974; B.S., Wildlife Biology,<br />

Michigan State University, 1981; M.S.,<br />

Wildlife Biology, Humboldt State<br />

University, 1985; Ph.D., Wildland<br />

Resource Science, University of Cali<strong>for</strong>nia,<br />

Berkeley, 1989.<br />

Current Position: Project Leader, Research<br />

Wildlife Biologist, USDA Forest Service,<br />

Rocky Mountain Forest and Range<br />

Experiment Station, Flagstaff, Arizona.<br />

Fernando Clemente, Wildlife Biologist.<br />

Education: B.S., Animal Science, University<br />

of Chapingo, Mexico, 1977; M.S.,<br />

Wild Animal Nutrition, Colegio De<br />

Postgraduados, Mexico, 1984; Ph.D.,<br />

Range and Wildlife Management, New<br />

Mexico State University, 1992.<br />

Current Position: Head, Department of Wildlife<br />

Science, and Wildlife Professor, Colegio<br />

De Postgraduados, Campus San Luis<br />

Potosi, Mexico.<br />

James L. Dick, Jr., Silviculturist.<br />

Education: B.S., Forestry, University of Montana,<br />

1967; M.S., Forest Resources,<br />

Pennsylvania State University, 1972.<br />

Current Position: Forester, Recreation Staff Unit,<br />

USDA Forest Service, Southwestern<br />

Region, Albuquerque, New Mexico.<br />

Alan B. Franklin, <strong>Spotted</strong> <strong>Owl</strong> Researcher.<br />

Education: B.S., Wildlife Biology, Cornell<br />

University, 1979; M.S., Wildlife Biology,<br />

Humboldt State University, 1987;<br />

Doctoral Candidate, Department of<br />

Fisheries and Wildlife, Colorado State<br />

University.<br />

Current Position: Project Leader, Humboldt<br />

Volume I/Appendix A<br />

APPENDIX APPENDIX A<br />

A<br />

163<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

State University Foundation, Humboldt<br />

State University, Arcata, Cali<strong>for</strong>nia.<br />

Joseph L. Ganey, <strong>Spotted</strong> <strong>Owl</strong> Researcher.<br />

Education: B.S., Wildlife Biology, Humboldt<br />

State University, 1981; M.S., Biology,<br />

Nor<strong>the</strong>rn Arizona University, 1988;<br />

Ph.D., Zoology, Nor<strong>the</strong>rn Arizona<br />

University, 1991.<br />

Current Position: Research Wildlife Biologist,<br />

USDA Forest Service, Rocky Mountain<br />

Forest and Range Experiment Station,<br />

Flagstaff, Arizona.<br />

W. H. Moir, Ecologist.<br />

Education: Ph.D., Botany and Soils,<br />

Washington State University, 1965.<br />

Current Position: Research Ecologist, USDA<br />

Forest Service, Rocky Mountain Forest<br />

and Range Experiment Station, Fort<br />

Collins, Colorado.<br />

Sarah E. Rinkevich, Wildlife Biologist.<br />

Education: B.S., Wildlife and Fisheries Science,<br />

University of Arizona, 1987; M.S.,<br />

Wildlife Biology, Humboldt State<br />

University, 1991.<br />

Current Position: Fish and Wildlife Biologist,<br />

USDI Fish and Wildlife Service, New<br />

Mexico Ecological Services State Office,<br />

Albuquerque, New Mexico.<br />

Dean L. Urban, Landscape Ecologist.<br />

Education: B.A., Botany and Zoology, Sou<strong>the</strong>rn<br />

Illinois University, 1978; M.A., Zoology<br />

(Wildlife Ecology), Sou<strong>the</strong>rn Illinois<br />

University, 1981; Ph.D., Ecology,<br />

University of Tennessee, 1986.<br />

Current Position: Assistant Professor, School of<br />

Environment, Duke University.<br />

James P. Ward, Jr., <strong>Spotted</strong> <strong>Owl</strong> Researcher.<br />

Education: B.S., Wildlife Biology, Humboldt<br />

State University, 1985; M.S., Natural<br />

Resources (wildlife science emphasis),<br />

Humboldt State University, 1990;<br />

Doctoral Candidate, Department of<br />

Biology, Colorado State University.


Current Position: Wildlife Biologist, USDA<br />

Forest Service, Rocky Mountain Forest<br />

and Range Experiment Station, Fort<br />

Collins, Colorado.<br />

Gary C. White, Population Ecologist.<br />

Education: B.S., Fisheries and Wildlife Biology,<br />

Iowa State University, 1970; M.S.,<br />

Wildlife Biology, University of Maine-<br />

Orono, 1972; Ph.D., Zoology, Ohio<br />

State University, 1976; Post Doctorate,<br />

Wildlife Biology, Utah State University,<br />

1976-77.<br />

Current Position: Professor, Department of<br />

Fishery and Wildlife, Colorado State<br />

University, Fort Collins, Colorado.<br />

RECOVERY RECOVERY TEAM-<br />

TEAM-<br />

FISH FISH AND AND WILDLIFE WILDLIFE SERVICE<br />

SERVICE<br />

LIAISON:<br />

LIAISON:<br />

Steven L. Spangle, Fish and Wildlife Biologist —<br />

Regional Listing Coordinator,<br />

USDI Fish and Wildlife Service, Southwestern<br />

Regional Office, Albuquerque,<br />

New Mexico.<br />

CONSULTANTS:<br />

CONSULTANTS:<br />

Pat Christgau, Coordinator, <strong>Mexican</strong> <strong>Spotted</strong><br />

<strong>Owl</strong> Management, Arizona Game and<br />

Fish Department, Phoenix, Arizona.<br />

Jack F. Cully, Jr., Assistant Unit Leader-Wildlife,<br />

National Biological Service, Kansas<br />

Cooperative Fish and Wildlife Research<br />

Unit, Kansas State University, Manhattan,<br />

Kansas.<br />

Frank P. Howe, Utah Partners in Flight<br />

Volume I/Appendix A<br />

164<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Coordinator, Utah Division of Wildlife<br />

Resources, Salt Lake City, Utah.<br />

Tim Keitt, Graduate Research Assistant,<br />

Department of Biology, University<br />

of New Mexico, Albuquerque, New<br />

Mexico<br />

Tom Spalding, Deputy Director, Arizona<br />

Department of Game and Fish,<br />

Phoenix, Arizona.<br />

Steve Thompson, Biological Technician, Forest<br />

Resources Program, San Carlos Apache<br />

Tribe, San Carlos, Arizona.<br />

Robert Vahle, Program Manager, Arizona<br />

Game and Fish Department,<br />

Region 1, Pinetop, Arizona.<br />

MEETING MEETING FACILITATOR:<br />

FACILITATOR:<br />

FACILITATOR:<br />

Kate W. Grandison, Intermountain Regional<br />

<strong>Spotted</strong> <strong>Owl</strong> Coordinator,<br />

Dixie National Forest, Cedar City, Utah.<br />

ADMINISTRATIVE ADMINISTRATIVE ASSISTANT:<br />

ASSISTANT:<br />

Brenda Witsell, Biological Technician, USDA<br />

Forest Service, Rocky Mountain Forest<br />

and Range Experiment Station, Flagstaff,<br />

Arizona.


March 29 to April 4, 1993<br />

Albuquerque, New Mexico<br />

April 27-30, 1993<br />

Flagstaff, Arizona<br />

May 17-21, 1993<br />

Sierra Vista, Arizona<br />

June 23-25, 1993<br />

Alamogordo, New Mexico<br />

July 12-16, 1993<br />

Flagstaff, Arizona<br />

August 16-19, 1993<br />

Pinetop, Arizona<br />

September 14-16, 1993<br />

Fort Collins, Colorado<br />

October 13-15, 1993<br />

Albuquerque, New Mexico<br />

January 10-14, 1994<br />

Flagstaff, Arizona<br />

February 22-25, 1994<br />

Fort Collins, Colorado<br />

March 14-16, 1994<br />

Albuquerque, New Mexico<br />

Volume I/Appendix B<br />

APPENDIX APPENDIX B<br />

B<br />

SCHEDULE SCHEDULE OF OF TEAM TEAM MEETINGS<br />

MEETINGS<br />

165<br />

April 25-29, 1994<br />

Aguascalientes, Mexico<br />

May 23-27, 1994<br />

Cedar City, Utah<br />

June 27 to July 1, 1994<br />

Flagstaff, Arizona<br />

August 8-12, 1994<br />

Flagstaff, Arizona<br />

September 7-9, 1994<br />

Fort Collins, Colorado<br />

September 29, 1994<br />

Albuquerque, New Mexico<br />

October 18-19, 1994<br />

Phoenix, Arizona<br />

February 13-17, 1995<br />

Phoenix, Arizona<br />

June 19-23, 1995<br />

Phoenix, Arizona<br />

July 17-21, 1995<br />

Flagstaff, Arizona<br />

August 14-18, 1995<br />

Albuquerque, New Mexico<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


APPENDIX APPENDIX C<br />

C<br />

SCHEDULE SCHEDULE OF OF FIELD FIELD VISITS<br />

VISITS<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

DATE DATE<br />

PLACE PLACE PLACE<br />

COORDINATOR<br />

COORDINATOR<br />

April 4, 1993 Walnut Canyon/Bar M Joe Ganey,<br />

Canyon, Coconino NF, Arizona Hea<strong>the</strong>r Green<br />

May 20, 1993 Huachuca, Santa Rita and, Russell Duncan<br />

Patagonia Mountains, Arizona Steve Spiech<br />

June 23, 1993 Sacramento Mountains, Danney Salas<br />

Lincoln NF, New Mexico Pat Ward<br />

August 8, 1993 Overflight, Gila NF, Bruce Anderson<br />

New Mexico Steve Servis<br />

August 20, 1993 Fort Apache Indian Reservation, Joe Jojola<br />

Arizona<br />

April 4, 1994 Sierra Fria, National Wildlife Council<br />

Aguascalientes, Mexico<br />

May 11-13, 1994 San Carlos Apache Indian Reservation, Steve Thompson<br />

Arizona Tim Wilhite<br />

May 5, 1994 Zion National Park, Utah Sarah Rinkevich<br />

Volume I/Appendix C<br />

166


A A REFERENCE REFERENCE FOR FOR ENGLISH<br />

ENGLISH<br />

AND AND LATIN LATIN NAMES<br />

NAMES<br />

English names were used in <strong>the</strong> text of<br />

this <strong>Recovery</strong> <strong>Plan</strong> to make smoo<strong>the</strong>r reading<br />

and to improve comprehension among people<br />

who are not familiar with Latin names. Names<br />

are arranged in alphabetical order of families <strong>for</strong><br />

plants because this is <strong>the</strong> conventional way of<br />

arranging <strong>the</strong>m in most commonly accessible<br />

tree and wildflower manuals. Species within <strong>the</strong><br />

families are listed alphabetically by Latin names.<br />

Animal names are listed phylogenetically, or<br />

taxonomically, because this system prevails in<br />

most commonly accessible books of birds and<br />

mammals, <strong>the</strong> principal species reported here.<br />

Family names are also provided <strong>for</strong> <strong>the</strong> animals<br />

as an aid <strong>for</strong> fur<strong>the</strong>r investigation.<br />

Aceraceae<br />

Aceraceae<br />

PLANTS<br />

PLANTS<br />

Canyon Acer<br />

(Bigtooth) maple grandidentatum<br />

Boxelder Acer negundo<br />

Chenopodiaceae<br />

Chenopodiaceae<br />

Shadscale Atriplex sp.<br />

Cupressaceae<br />

Cupressaceae<br />

Arizona cypress Cupressus arizonica<br />

Juniper Juniperus sp.<br />

Ericaceae<br />

Ericaceae<br />

Madrone Arbutus sp.<br />

Manzanita Arctostaphylos sp.<br />

Fabaceae<br />

Fabaceae<br />

Mesquite Prosopis sp.<br />

Volume I/Appendix D<br />

APPENDIX APPENDIX D<br />

D<br />

167<br />

Fagaceae<br />

Fagaceae<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Arizona white oak Quercus arizonica<br />

Quercus chihuahuensis<br />

Quercus coccolobifolia<br />

Emory oak Quercus emoryi<br />

Gambel oak Quercus gambelii<br />

Quercus gentryi<br />

Gray oak Quercus grisea<br />

Silverleaf oak Quercus hypoleucoides<br />

Quercus laeta<br />

Quercus potosina<br />

Quercus resinosa<br />

Netleaf oak Quercus rugosa<br />

Wavyleaf oak Quercus undulata<br />

Papilionoideae<br />

Papilionoideae<br />

Papilionoideae<br />

New Mexico locust Robinia neomexicana<br />

Pinaceae<br />

Pinaceae<br />

White fir Abies concolor<br />

Blue Spruce Picea pungens<br />

Pinyon pine Pinus edulis<br />

Limber pine Pinus flexilis<br />

Western white pine Pinus monticola<br />

Ponderosa pine Pinus ponderosa<br />

Aztec pine Pinus teocote<br />

Southwestern Pinus strobi<strong>for</strong>mis<br />

white pine<br />

Douglas-fir Pseudotsuga menziesii<br />

Redwood Sequoia sempervirens<br />

Platanaceae<br />

Platanaceae<br />

Arizona Sycamore Platanus wrightii<br />

Salicaceae<br />

Salicaceae<br />

Narrowleaf cottonwood Populus angustifolia<br />

Trembling aspen Populus tremuloides<br />

Viscaceae<br />

Viscaceae<br />

Dwarf mistletoe Arceuthobium sp.<br />

Zygophyllaceae<br />

Zygophyllaceae<br />

Creosotebush Larrea sp.


Tortricidae<br />

Tortricidae<br />

ANIMALS<br />

ANIMALS<br />

INVERTEBRATES<br />

INVERTEBRATES<br />

Western Choristoneura<br />

spruce budworm <strong>occidentalis</strong><br />

Scolytidae<br />

Scolytidae<br />

Round-headed beetle Dendroctonus<br />

adjunctus<br />

Danaidae<br />

Danaidae<br />

Monarch Butterfly Danaus plexippus<br />

Birds<br />

Birds<br />

Accipitridae<br />

Accipitridae<br />

VERTEBRATES<br />

VERTEBRATES<br />

VERTEBRATES<br />

Golden eagle Aquila chrysaetos<br />

Nor<strong>the</strong>rn goshawk Accipter gentilis<br />

Red-tailed hawk Buteo jamaicensis<br />

Strigidae Strigidae<br />

Strigidae<br />

Great horned owl Bubo virginianus<br />

<strong>Spotted</strong> owl <strong>Strix</strong> <strong>occidentalis</strong><br />

Cali<strong>for</strong>nia spotted owl S. o. <strong>occidentalis</strong><br />

<strong>Mexican</strong> spotted owl S. o. <strong>lucida</strong><br />

Nor<strong>the</strong>rn spotted owl S. o. caurina<br />

Barred owl <strong>Strix</strong> varia<br />

Fulvous owl <strong>Strix</strong> fulvescens<br />

Tawny owl <strong>Strix</strong> aluco<br />

Psittacidae<br />

Psittacidae<br />

Yellow-headed parrot Amazona ochrocephala<br />

Mammals<br />

Mammals<br />

Soricidae<br />

Soricidae<br />

Masked shrew Sorex cinereus<br />

Vagrant shrew Sorex vagrans<br />

Volume I/Appendix D<br />

168<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Montane shrew Sorex monticouls<br />

Dusky shrew Sorex obscurus<br />

Water shrew Sorex palustris<br />

Leporidae<br />

Leporidae<br />

Desert cottontail Sylvilagus audubonii<br />

Eastern cottontail Sylvilagus floridanus<br />

Nuttall’s cottontail Sylvilagus nuttallii<br />

Sciuridae<br />

Sciuridae<br />

Nor<strong>the</strong>rn Glaucomys sabrina<br />

flying squirrel<br />

Golden-mantled Spermophilus<br />

ground squirrel lateralis<br />

Rock squirrel Spermophilus<br />

variegatus<br />

Gray-collared Tamias cinereicollus<br />

chipmunk<br />

Gray-footed chipmunk Tamias canipes<br />

Least chipmunk Tamias minimus<br />

Cliff chipmunk Tamias dorsalis<br />

Red squirrel Tamiasciurus<br />

hudsonicus<br />

Geomyidae Geomyidae<br />

Geomyidae<br />

Botta’s pocket gopher Thomomys bottae<br />

Sou<strong>the</strong>rn pocket Thomomys umbrinus<br />

gopher<br />

Nor<strong>the</strong>rn pocket Thomomys talpoides<br />

gopher<br />

Heteromyidae<br />

Heteromyidae<br />

Great Basin Perognathus parvus<br />

pocket mouse<br />

Muridae<br />

Muridae<br />

Pinyon mouse Peromyscus truei<br />

Brush mouse Peromyscus boylei<br />

Canyon mouse Peromyscus crinitis<br />

Rock mouse Peromyscus difficilis<br />

Deer mouse Peromyscus<br />

maniculatus<br />

White-footed mouse Peromyscus leucopus<br />

Bushy-tailed woodrat Neotoma cinerea<br />

Desert woodrat Neotoma lepida


<strong>Mexican</strong> woodrat Neotoma mexicana<br />

Stephens woodrat Neotoma stephensi<br />

White-throated Neotoma albigula<br />

woodrat<br />

Long-tailed vole Microtus longicaudus<br />

Meadow vole Microtus<br />

pennsylvanicus<br />

<strong>Mexican</strong> vole Microtus mexicanus<br />

Montane vole Microtus montanus<br />

Dipodidae<br />

Dipodidae<br />

Meadow Zapus hudsonius<br />

jumping mouse<br />

Western Zapus princeps<br />

jumping mouse<br />

Volume I/Appendix D<br />

169<br />

Mustelidae<br />

Mustelidae<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Long-tailed weasel Mustela frenata<br />

Trichechidae<br />

Trichechidae<br />

Manatee Trichechus manatus


APPENDIX APPENDIX E<br />

E<br />

AGENCIES AGENCIES AND AND PERSONS PERSONS COMMENTING<br />

COMMENTING<br />

ON ON ON DRAFT DRAFT RECOVERY RECOVERY PLAN<br />

PLAN<br />

Critical review of planning documents by<br />

those that must implement <strong>the</strong>m and o<strong>the</strong>rs<br />

with relevant expertise is essential to producing<br />

management plans that are scientifically credible<br />

and feasible to implement. This appendix lists<br />

agencies and persons who reviewed <strong>the</strong> draft<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong> and provided<br />

comments on <strong>the</strong> document. These<br />

comments are in large part responsible <strong>for</strong><br />

production of a final <strong>Recovery</strong> <strong>Plan</strong> that <strong>the</strong><br />

<strong>Recovery</strong> Team believes is much improved over<br />

<strong>the</strong> draft version. Many comments were directly<br />

responsible <strong>for</strong> <strong>Recovery</strong> <strong>Plan</strong> revision, while<br />

many comments that were not incorporated<br />

provoked considerable thought and lively discus-<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

sion during <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>-revision period.<br />

The <strong>Recovery</strong> Team is grateful to those who<br />

spent valuable time contributing to this final<br />

<strong>Recovery</strong> <strong>Plan</strong>.<br />

PEER PEER PEER REVIEW<br />

REVIEW<br />

The following persons were specifically asked<br />

to review <strong>the</strong> draft <strong>Recovery</strong> <strong>Plan</strong> or portions<br />

<strong>the</strong>reof, as indicated. “Parts Reviewed” referred<br />

to below relates to <strong>the</strong> draft <strong>Recovery</strong> <strong>Plan</strong>, not<br />

this document. Each persons’ affiliation is listed.<br />

In addition, scientific and professional organizations<br />

that requested an individual’s review are so<br />

indicated.<br />

Commentor Commentor<br />

Par ar arts arts<br />

ts<br />

Review eview eview eviewed eview ed Affiliation Affiliation<br />

Organization rganization<br />

Forsman, E.D. All Pacific Northwest Research American Ornithologists’<br />

Station, U.S. Forest Service Union<br />

Corvallis, OR<br />

Gutiérrez, R.J. I, II Humboldt State University<br />

Arcata, CA<br />

Holthausen, R. I,II, III U.S. Forest Service<br />

Corvallis, OR<br />

King, R. II, III Rocky Mountain Forest and<br />

Range Experiment Station<br />

U.S. Forest Service<br />

Fort Collins, CO<br />

LaHaye, W. II.F, II.G Humboldt State University,<br />

Arcata, CA<br />

Meslow, E.C. All Wildlife Management Institute The Wildlife Society;<br />

Corvallis, OR Wildlife Management<br />

Institute<br />

Morrison, M.L. II.G, II.H University of Arizona<br />

Tucson, AZ<br />

Volume I/Appendix E<br />

170


Commentor Commentor<br />

Par ar arts ar ts<br />

Review eview eview eviewed eview ed Affiliation Affiliation<br />

Organization rganization<br />

Pollock, K.H. II.E, III.A, North Carolina State University<br />

III.D Raleigh, NC<br />

Raphael, M.G. I, II, III Pacific Northwest Research<br />

Station, U.S. Forest Service<br />

Olympia, WA<br />

Stacey, P. II.F, II.G, University of Nevada<br />

III Reno, NV<br />

Verner, J. II.G, III Pacific Southwest Research<br />

Station, U.S. Forest Service<br />

Fresno, CA<br />

AGENCY AGENCY AGENCY REVIEW REVIEW<br />

REVIEW<br />

The following agencies provided comments<br />

on <strong>the</strong> draft <strong>Recovery</strong> <strong>Plan</strong>. Federal agencies are<br />

listed first, followed by State and County agencies.<br />

Agency name is followed by signator; o<strong>the</strong>r<br />

reviewers are listed in ( ) when identified by <strong>the</strong><br />

agency.<br />

National Park Service, Sou<strong>the</strong>rn Arizona Group;<br />

Jerry Belson, General Superintendent.<br />

(Benson, L.).<br />

U.S. Bureau of Indian Affairs, Albuquerque Area<br />

Office; Patrick A. Hayes, Area Director.<br />

(Schwab, B.).<br />

U.S. Bureau of Indian Affairs, Sou<strong>the</strong>rn Ute<br />

Agency; Charles A. Recker, Superintendent.<br />

(Friedley, J.; Recker, T.).<br />

U.S. Bureau of Land Management, Utah; Mat<br />

Millenbach, State Director.<br />

(Stringer, W.).<br />

U.S. Fish and Wildlife Service, Arizona Ecological<br />

Services State Office; Sam F. Spiller,<br />

State Supervisor. (James, M.; Muiznieks,<br />

B.; Palmer, B.).<br />

Volume I/Appendix E<br />

171<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

U.S. Fish and Wildlife Service, New Mexico<br />

Ecological Services State Office; Jennifer<br />

Fowler-Propst, State Supervisor.<br />

(Torres, C.).<br />

U.S. Fish and Wildlife Service, Mountain-Prairie<br />

Region; James M. Lutey, Acting Assistant<br />

Regional Director, Ecological Services.<br />

U.S. Forest Service, Southwestern Region;<br />

Charles W. Cartwright, Jr., Regional<br />

Forester. (Beyerhelm, C.; Birkland,<br />

C.; Briggs, A.; Brown, A.; Casper, L.;<br />

Cassidy, R.; Dargan, C.; Deaver, R.;<br />

DeLorenzo, D.; Derby, J.; Ellenwood, J.;<br />

Ewers, S.; Fletcher, R.; Gerritsma, J.;<br />

Green, H.; Herron, M.; Higgins, B.;<br />

Holbrook, C.; Hollis, H.; Holmstrom,<br />

D.; Johnson, D.; Kill, D.; Lucero, L.;<br />

MacIvor, J.; Madril, A.; Man<strong>the</strong>i, M.;<br />

Martinez, J.; Menasco, K.; Nelson, J.;<br />

Randall-Parker, T.; Rethlake, K.; Rolf, J.;<br />

Schaal, L.; Shafer, J.; Sheppard, G.;<br />

Spoerl, P.; Stahn, R.; Taylor, C,; Vigil,<br />

G.; Wistrand, H.; Zumwalt, M.).<br />

U.S. Forest Service, Intermountain Region; Dale<br />

N. Bosworth, Regional Forester. (Botts,<br />

J.; Egnew, A.; Grandison, K.; Gray, S.;<br />

Hayman, R.).


U.S. Forest Service, Rocky Mountain Region;<br />

Elizabeth Estill, Regional Forester.<br />

(Player, R.).<br />

Arizona Game and Fish Department; Duane L.<br />

Shroufe, Director. (Johnson, T.).<br />

Arizona State Land Department; M. Jean<br />

Hassell, Commissioner.<br />

Otero County Commission; Richard L. Zierlein,<br />

Chairman.<br />

San Juan County Commission;<br />

Ty Lewis, Chairman.<br />

OTHER OTHER GROUPS<br />

GROUPS<br />

AND AND AND INDIVIDUALS<br />

INDIVIDUALS<br />

INDIVIDUALS<br />

Applied Ecosystem Management, Flagstaff, AZ;<br />

Tod Hull. (On behalf of: Nor<strong>the</strong>rn<br />

Arizona Loggers Association; Precision<br />

Pine and Timber Company; Reidhead<br />

Bro<strong>the</strong>rs Lumber Company; Stone<br />

Forest Industries).<br />

Defenders of Wildlife, Washington, DC; Robert<br />

M. Ferris, Director, Species Conservation<br />

Division; Gregory J. Sater, Wildlife<br />

Counsel, Legal Division.<br />

Forest Conservation Council, Santa Fe, NM;<br />

John Talberth, Executive Director.<br />

Volume I/Appendix E<br />

172<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Southwest Center <strong>for</strong> Biological Diversity, Silver<br />

City, NM; Kieran Suckling, Executive<br />

Director. (Attachment signed by Dennis<br />

Morgan, Research Associate, Southwest<br />

Center <strong>for</strong> Biological Diversity; Kieran<br />

Suckling, Executive Director, Southwest<br />

Center <strong>for</strong> Biological Diversity; Sharon<br />

Galbreath, Chairperson, Grand Canyon<br />

Chapter, Sierra Club; Samuel Hitt,<br />

Director, Forest Guardians; Joanie Berde,<br />

Carson Forest Watch; Tom Ribe, Public<br />

Forestry Foundation; Tom H. Wootten,<br />

Conservation Chair, Mesilla Valley<br />

Audubon Society; Mary Lou Jones, Zuni<br />

Mountain Coalition; Joseph Feller; Dave<br />

Henderson, Southwest Forest Alliance;<br />

Charles Babbitt, President, Maricopa<br />

Audubon Society; John Talberth, Executive<br />

Director, Forest Conservation<br />

Council; Jim Powers, Prescott National<br />

Forest Friends; Gary Simpson, Nor<strong>the</strong>rn<br />

New Mexico Chapter, Wilderness<br />

Watch; Eleanor G. Wooten, Vice President,<br />

T & E, Inc.; Gwen Wardwell,<br />

Director, Rio Grande Chapter, Sierra<br />

Club; Mike Siedman; Jeff Burgess.).<br />

Stone Forest Industries, Flagstaff, AZ; Steve C.<br />

Bennett, Regional Manager.<br />

Mark Herron, Santa Fe, NM.<br />

Terry Johnson, Los Alamos, NM.<br />

Dennis R. Kingsbury, Munds Park, AZ.


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Supporting Documents<br />

Volume II


MEXICAN MEXICAN SPOTTED SPOTTED OWL OWL RECOVERY RECOVERY PLAN<br />

PLAN<br />

Volume Volume Volume II II - - Technical Technical Supporting Supporting In<strong>for</strong>mation<br />

In<strong>for</strong>mation<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

This volume consists of chapters on aspects of <strong>Mexican</strong> spotted owl natural history that were<br />

developed during preparation of <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong> (<strong>Recovery</strong> <strong>Plan</strong>). These<br />

treatises provide much of <strong>the</strong> in<strong>for</strong>mation upon which <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>, especially <strong>the</strong> recovery<br />

recommendations in Part III of Volume I, are based. Much of <strong>the</strong> material in Volume II is highly<br />

technical in nature and, although important, was not considered appropriate <strong>for</strong> inclusion in a working<br />

implementation document like Volume I of <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>. The <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

Team and <strong>the</strong> Fish and Wildlife Service believe that <strong>the</strong>se technical papers should, however, be available<br />

to anyone who would like a detailed account of subject matter contained herein.<br />

Since <strong>the</strong> material detailed in Volume II was integral in developing <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong>, <strong>the</strong> salient<br />

points from each of <strong>the</strong> Volume II chapters are summarized in Part II of Volume I. This makes<br />

Volume I a stand-alone document containing <strong>the</strong> most relevant in<strong>for</strong>mation needed to implement <strong>the</strong><br />

<strong>Recovery</strong> <strong>Plan</strong> and understand <strong>the</strong> reasons behind <strong>the</strong> management recommendations contained<br />

<strong>the</strong>rein.<br />

Citations of material contained in this volume should read as follows:<br />

[Author(s)] 1995. Pages [ - ] in USDI Fish and Wildlife Service. <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

<strong>Plan</strong>, Volume II.<br />

CONTENTS<br />

CONTENTS<br />

Chapter Chapter 1: 1: D DDistr<br />

D istr istribution istr ibution and and A AAbundance<br />

A bundance<br />

James P. Ward, Jr., Alan B. Franklin, Sarah E. Rinkevich, and Fernando Clemente............... 14 pages<br />

Chapter Chapter 2: 2: P PPopulation<br />

P opulation B BBiolog<br />

BB<br />

iolog iology iolog<br />

Gary C. White, Alan B. Franklin, and James P. Ward, Jr...................................................... 25 pages<br />

Chapter Chapter 3: 3: Landscape Landscape Analysis Analysis and and M MMetapopulation<br />

M etapopulation S SStr<br />

S tr tructur tr uctur ucture uctur<br />

Tim Keitt, Alan B. Franklin, and Dean Urban ................................................................... 16 pages<br />

Chapter Chapter Chapter 4: 4: H HHabitat<br />

H abitat R RRelationships<br />

R elationships<br />

Joseph L. Ganey and James L. Dick, Jr. ............................................................................... 42 pages<br />

Chapter Chapter 5: 5: P PPrey<br />

P ey E EEcolog<br />

E Ecolog<br />

colog cology colog<br />

James P. Ward, Jr. and William M. Block ............................................................................ 48 pages<br />

Volume II i


Chapter Chapter 1<br />

1<br />

Volume II/List of Tables<br />

List List of of Tables<br />

Tables<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able 1.1. 1.1. Historical records and minimum numbers of <strong>Mexican</strong> spotted<br />

owls found during planned surveys .................................................................................................2-3<br />

Chapter Chapter 2<br />

2<br />

Table able 2.1. 2.1. 2.1. Time periods and number of capture histories from three<br />

study areas used in estimating <strong>Mexican</strong> spotted owl survival. ............................................................. 5<br />

Table able 2.2. 2.2. Apparent survival and recapture probability from three<br />

study areas used in estimating <strong>Mexican</strong> spotted owl survival .............................................................. 6<br />

Table able 2.3. 2.3. Description of radio-telemetry studies conducted on adult<br />

and subadult <strong>Mexican</strong> spotted owls ................................................................................................... 8<br />

Table able 2.4. 2.4. Estimates of true survival <strong>for</strong> <strong>Mexican</strong> spotted owls<br />

based on radio-telemetry data. .............................................................................................. 9<br />

Table able 2.5. 2.5. Number of management territories checked one or more<br />

times during <strong>the</strong> <strong>Mexican</strong> spotted owl monitoring program of FS Region 3 ....................................10<br />

Table able 2.6. 2.6. Persistence of a <strong>Mexican</strong> spotted owl pair on a territory<br />

<strong>for</strong> <strong>for</strong>mal and in<strong>for</strong>mal monitoring data ........................................................................................10<br />

Table able 2.7. 2.7. Persistence of a pair of <strong>Mexican</strong> spotted owls on a territory<br />

<strong>for</strong> <strong>the</strong> five recovery units contained in <strong>the</strong> FS Region 3 monitoring data base ................................11<br />

Table able 2.8. 2.8. Estimates of number of young fledged/pair and fecundity<br />

on two demographic study areas .....................................................................................................13<br />

Table able 2.9. 2.9. Mean number of young fledged from 1989-1993,<br />

based on FS monitoring data per pair .............................................................................................14<br />

Table able able 2.10. 2.10. Mean number of young produced/pair from 1989-1993,<br />

based on monitoring data ...............................................................................................................15<br />

Table able 2.11. 2.11. Mean number of young produced per territory from 1989-1993,<br />

based on monitoring data ...............................................................................................................15<br />

Table able 2.12. 2.12. Parameters used to estimate l <strong>for</strong> <strong>Mexican</strong> spotted<br />

owls on <strong>the</strong> GSA and CSA study areas .............................................................................................20<br />

Table able 2.13. 2.13. 2.13. Estimates and standard errors of l <strong>for</strong> female <strong>Mexican</strong><br />

spotted owls on <strong>the</strong> CSA and GSA study areas ................................................................................20<br />

ii


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able 2.14. 2.14. Estimates of population parameters <strong>for</strong> <strong>Mexican</strong><br />

spotted owls on <strong>the</strong> GSA and CSA study areas ................................................................................21<br />

Table able 2.15. 2.15. Number of <strong>Mexican</strong> spotted owl territories checked<br />

and <strong>the</strong> percent of <strong>the</strong>m occupied by a pair of owls <strong>for</strong> 1989-1993..................................................22<br />

Chapter Chapter 4<br />

4<br />

Table able 4.1. 4.1. Percent of nest and roost sites in various vegetation<br />

types in different <strong>Recovery</strong> Units ...................................................................................................... 4<br />

Table able 4.2. 4.2. Home-range sizes of radio-marked <strong>Mexican</strong> spotted owls............................................ 6-7<br />

Table able 4.3. 4.3. Size of nocturnal activity centers of radio-tagged pairs of<br />

<strong>Mexican</strong> spotted owls in <strong>the</strong> Upper Gila Mountains and Basin and<br />

Range-East <strong>Recovery</strong> Units ............................................................................................................... 8<br />

Table able able 4.4. 4.4. Use of <strong>for</strong>est types <strong>for</strong> <strong>for</strong>aging by radio-tagged <strong>Mexican</strong> spotted<br />

owls on three study areas in nor<strong>the</strong>rn Arizona. ................................................................................10<br />

Table able able 4.5. 4.5. 4.5. Selected characteristics of nest stands of <strong>Mexican</strong> spotted owls in<br />

<strong>the</strong> Upper Gila Mountains and Basin and Range-East <strong>Recovery</strong> Units.............................................12<br />

Table able 4.6. 4.6. Habitat characteristics sampled at <strong>Mexican</strong> spotted owl nest sites<br />

on three Ranger districts, Apache-Sitgreaves National Forest, Upper Gila<br />

Mountains <strong>Recovery</strong> Unit...............................................................................................................14<br />

Table able 4.7. 4.7. Habitat characteristics at <strong>Mexican</strong> spotted owl nest and<br />

randomly located sites in <strong>the</strong> Tularosa Mountains, New Mexico; Upper<br />

Gila Mountains RU. .......................................................................................................................15<br />

Table able 4.8a. 4.8a. Habitat characteristics of <strong>Mexican</strong> spotted owl nest sites<br />

in <strong>the</strong> Upper Gila Mountains <strong>Recovery</strong> Unit...................................................................................20<br />

Table able 4.8b. 4.8b. 4.8b. Habitat characteristics of <strong>Mexican</strong> spotted owl nest sites<br />

in <strong>the</strong> Basin and Range-East <strong>Recovery</strong> Unit ....................................................................................21<br />

Table able 4.9. 4.9. 4.9. Habitat characteristics on circular plots within home ranges<br />

of radio-tagged <strong>Mexican</strong> spotted owls .............................................................................................23<br />

Table able 4.10. 4.10. Habitat characteristics at <strong>Mexican</strong> spotted owl roost and<br />

random sites in <strong>the</strong> Tularosa Mountains, New Mexico; Upper Gila<br />

Mountains RU ...............................................................................................................................24<br />

Table able 4.11a. 4.11a. Selected characteristics of roost sites used by radio-tagged<br />

<strong>Mexican</strong> spotted owls in <strong>the</strong> Colorado Plateau <strong>Recovery</strong> Unit. ........................................................27<br />

Table able 4.11b. 4.11b. 4.11b. Selected characteristics of roost sites used by radio-tagged<br />

<strong>Mexican</strong> spotted owls in <strong>the</strong> Upper Gila Mountains <strong>Recovery</strong> Unit ................................................. 28<br />

Volume II/List of Tables iii


iv<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able 4.11c. 4.11c. Selected characteristics of roost sites used by radio-tagged<br />

<strong>Mexican</strong> spotted owls in <strong>the</strong> Basin and East-Range <strong>Recovery</strong> Unit ..................................................29<br />

Table able 4.12. 4.12. Seasonal roost site characteristics of radio-tagged <strong>Mexican</strong><br />

spotted owls in ponderosa pine-Gambel oak <strong>for</strong>est, Arizona (Upper<br />

Gila Mountains <strong>Recovery</strong> Unit) ...................................................................................................... 31<br />

Table able 4.13. 4.13. Characteristics of roost sites used by two migrant <strong>Mexican</strong><br />

spotted owls on <strong>the</strong>ir winter range ..................................................................................................32<br />

Chapter Chapter 5<br />

5<br />

Table able 5.1. 5.1. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong><br />

spotted owls occurring in <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> subspecies' range ............................................ 3<br />

Table able 5.2. 5.2. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong><br />

spotted owls occurring in <strong>the</strong> central portion of <strong>the</strong> subspecies' range ............................................... 4<br />

Table able 5.3. 5.3. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong><br />

spotted owls occurring in <strong>the</strong> sou<strong>the</strong>rn portion of <strong>the</strong> subspecies' range ............................................ 5<br />

Table able 5.4. 5.4. 5.4. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls<br />

occurring in <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> subspecies' range ................................................................ 6<br />

Table able able 5.5. 5.5. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls<br />

occurring in <strong>the</strong> central portion of <strong>the</strong> subspecies' range ................................................................... 7<br />

Table able able 5.6. 5.6. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls<br />

occurring in <strong>the</strong> sou<strong>the</strong>rn portion of <strong>the</strong> subspecies' range ................................................................ 8<br />

Table able 5.7. 5.7. Animal species consumed by <strong>Mexican</strong> spotted owls in<br />

18 geographic areas ........................................................................................................................10<br />

Table able 5.8. 5.8. Factors influencing trends observed in <strong>Mexican</strong> spotted<br />

owl diets................................................................................................................................... 14-15<br />

Table able 5.9. 5.9. Factors influencing production of <strong>Mexican</strong> spotted owls<br />

in nor<strong>the</strong>rn Arizona, <strong>the</strong> Sacramento Mountains, New Mexico, and <strong>the</strong><br />

Tularosa Mountains, New Mexico, during 1991, 1992, and 1993 ............................................. 17-18<br />

Table able 5.10. 5.10. Prey comprising ³10% of relative frequency or biomass<br />

in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls.................................................................................................21<br />

Table able 5.11. 5.11. Descriptive statistics of selected habitat variables<br />

characterizing habitats of common mammalian prey of <strong>Mexican</strong> spotted owls ........................... 32-33<br />

Table able 5.12. 5.12. Habitat characteristics used by deer mice in three<br />

vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico ............................................ 34<br />

Volume II/List of Tables


<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table able 5.13. 5.13. 5.13. Habitat characteristics used by <strong>Mexican</strong> woodrats in two<br />

vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico ..............................................35<br />

Table able 5.14. 5.14. Habitat characteristics used by long-tailed voles in two<br />

vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico ..............................................36<br />

Table able 5.15. 5.15. Habitat characteristics used by <strong>Mexican</strong> voles in three<br />

vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico ..............................................37<br />

Table able 5.16. 5.16. 5.16. Habitat characteristics used by brush mice in xeric <strong>for</strong>ests<br />

of <strong>the</strong> Sacramento Mountains, New Mexico....................................................................................39<br />

Volume II/List of Tables v


Chapter Chapter 1<br />

1<br />

List List List of of Figures<br />

Figures<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e 1.1. 1.1. <strong>Recovery</strong> unit boundaries and current distribution of <strong>Mexican</strong><br />

spotted owls in <strong>the</strong> United States ..................................................................................................... 8<br />

Figur igur igure igur e 1.2. 1.2. Current distribution of <strong>Mexican</strong> spotted owls in Mexico ............................................... 9<br />

Figur igur igure igur e 1.3. 1.3. Density of (a) (a) all three subspecies compared among regions<br />

and (b) (b) (b) <strong>Mexican</strong> spotted owls among three <strong>for</strong>est types in <strong>the</strong> Sacramento<br />

Mountains, New Mexico ................................................................................................................11<br />

Chapter Chapter 2<br />

2<br />

Figur igur igure igur e 2.1. 2.1. Location of Coconino, Gila, and Sky Islands <strong>Mexican</strong> spotted<br />

owl population study areas in Arizona and New Mexico ................................................................... 5<br />

Figur igur igur igure igur e 2.2. 2.2. 2.2. Changes in occupancy of <strong>for</strong>mal and in<strong>for</strong>mal monitoring territories<br />

in <strong>the</strong> FS Region 3 monitoring database .........................................................................................23<br />

Chapter Chapter 3<br />

3<br />

Figur igur igure igur e 3.1. 3.1. Dispersal-distance relationship <strong>for</strong> radio-marked juvenile spotted owls ......................... 4<br />

Figur igur igure igur e 3.2. 3.2. The relationship between correlation length and minimum<br />

joining distance. .............................................................................................................................. 8<br />

Figur igur igure igur e 3.3. 3.3. Landscape mosaics of discrete clusters of Douglas-fir and ponderosa<br />

pine habitat types. ........................................................................................................................... 9<br />

Figur igur igure igur e 3.4. 3.4. Change in correlation length due to cluster removal as a function<br />

of distance ......................................................................................................................................10<br />

Figur igur igur igure igure<br />

e 3.5. 3.5. Maps with clusters colored to illustrate <strong>the</strong>ir rank importance,<br />

weighted by (uncorrected <strong>for</strong>) patch area. .......................................................................................11<br />

Figur igur igure igur e 3.6. 3.6. Patch importance to overall connectedness, normalized <strong>for</strong> patch area. ........................12<br />

Chapter Chapter 4<br />

4<br />

Figur igur igure igur e 4.1. 4.1. Diameter distributions of live trees sampled in nest stands in<br />

<strong>the</strong> Upper Gila Mountains and Basin and Range - East <strong>Recovery</strong> Units ...........................................11<br />

Figur igur igure igur e 4.2a. 4.2a. 4.2a. Tree species composition within nest- and random-stand<br />

plots, Upper Gila Mountains <strong>Recovery</strong> Unit ...................................................................................17<br />

Figur igur igur igure igure<br />

e 4.2b. 4.2b. 4.2b. Tree species composition within nest- and random-stand<br />

plots, Basin and Range-East <strong>Recovery</strong> Unit .....................................................................................18<br />

Volume II/List of Figures<br />

vi


Volume II/List of Figures vii<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figur igur igure igur e 4.3. 4.3. Diameter distributions of live trees sampled on nest,<br />

nest stand, and random stand plots ................................................................................................. 19<br />

Chapter Chapter 5<br />

5<br />

Figur igur igure igur e 5.1. 5.1. Cumulative distributions of prey in <strong>the</strong> diet of<br />

<strong>Mexican</strong> spotted owls .....................................................................................................................11<br />

Figur igur igure igure<br />

e 5.2. 5.2. Geographic variability in <strong>the</strong> food habits of<br />

<strong>Mexican</strong> spotted owls .....................................................................................................................12<br />

Figur igur igure igur e 5.3. 5.3. Reproductive success of <strong>Mexican</strong> spotted owls in <strong>the</strong><br />

Sacramento Mountains, New Mexico, nor<strong>the</strong>rn Arizona, and Tularosa<br />

Mountains, New Mexico ................................................................................................................19<br />

Figur igur igure igure<br />

e 5.4. 5.4. Biomass of (a) (a) common prey occurring in<br />

mixed-conifer <strong>for</strong>ests, (b) (b) frequencies of peromyscid mice consumed<br />

by <strong>Mexican</strong> spotted owls, and (c) (c) average number of owl young produced ......................................20<br />

Figur igur igure igure<br />

e 5.5. 5.5. 5.5. Average biomass of common prey of <strong>the</strong> <strong>Mexican</strong> spotted owl..................................... 26<br />

Figur igur igure igur e 5.6. 5.6. Trends in average density of common prey<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl ............................................................................................................. 29


Volume II/Chapter 1<br />

CHAPTER CHAPTER 1: 1: Distribution Distribution and and Abundance Abundance<br />

Abundance<br />

of of <strong>Mexican</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong>s<br />

<strong>Owl</strong>s<br />

James P. Ward, Jr., Alan B. Franklin, Sarah E. Rinkevich,<br />

and Fernando Clemente<br />

Knowledge of <strong>the</strong> distribution and abundance<br />

of <strong>Mexican</strong> spotted owls can provide<br />

insight into <strong>the</strong> subspecies’ geographic limits and<br />

habitat requirements. For example, standardized<br />

surveys <strong>for</strong> nor<strong>the</strong>rn spotted owls among different<br />

habitats have provided evidence of <strong>the</strong> owl’s<br />

affinity <strong>for</strong> older, densely layered <strong>for</strong>ests<br />

(Forsman et al. 1977; 1987, Thomas et al. 1990,<br />

Blakesley et al. 1992). In addition, distribution<br />

and abundance patterns often provide a foundation<br />

<strong>for</strong> more intensive natural and life history<br />

studies.<br />

For this recovery plan, we ga<strong>the</strong>red and<br />

examined in<strong>for</strong>mation on <strong>the</strong> distribution and<br />

abundance of <strong>Mexican</strong> spotted owls accumulated<br />

through 1993. We used this in<strong>for</strong>mation to (1)<br />

document historical and current extent of this<br />

subspecies, (2) help <strong>for</strong>mulate recovery unit<br />

boundaries, and (3) provide a template <strong>for</strong><br />

landscape-scale analyses.<br />

SOURCES SOURCES OF OF INFORMATION<br />

INFORMATION<br />

INFORMATION<br />

The quality and quantity of in<strong>for</strong>mation<br />

regarding <strong>the</strong> distribution and abundance<br />

<strong>Mexican</strong> spotted owls varies by source. Historical<br />

accounts exist from museum collections and<br />

anecdotal observations by early natural historians<br />

from throughout <strong>the</strong> owl’s range (reviewed in<br />

McDonald et al. 1991). These early observations<br />

are useful <strong>for</strong> documenting <strong>the</strong> owl’s known<br />

historical range. However, haphazard and frequently<br />

unknown methods by which <strong>the</strong> historical<br />

in<strong>for</strong>mation was obtained confound any<br />

attempt to infer change in <strong>the</strong> owl’s abundance<br />

from historical to present time. Modern accounts<br />

exist from incidental observations provided<br />

by amateur and professional biologists and<br />

from organized surveys conducted by natural<br />

resource management or research personnel.<br />

Incidental observations are similar in quality to<br />

historical accounts, frequently lacking sufficient<br />

in<strong>for</strong>mation <strong>for</strong> estimating population parameters<br />

or testing empirical hypo<strong>the</strong>ses. However,<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

when combined with results of planned surveys<br />

incidental observations can be used to document<br />

<strong>the</strong> current extent of <strong>the</strong> subspecies’ range.<br />

Results from planned surveys and demographic<br />

studies have provided <strong>the</strong> best available data on<br />

<strong>the</strong> owl’s abundance.<br />

<strong>Plan</strong>ned surveys <strong>for</strong> <strong>Mexican</strong> spotted owls in<br />

<strong>the</strong> United States have been conducted by landmanagement<br />

agencies since 1989 and by researchers<br />

in Mexico since 1992. Survey protocols<br />

were reviewed in 1990 and a more <strong>for</strong>mal<br />

program was subsequently developed to locate<br />

<strong>Mexican</strong> spotted owls (USDA Forest Service<br />

1990). <strong>Owl</strong> demographic studies began in <strong>the</strong><br />

Sky Island Mountains of Arizona in 1990<br />

(Duncan et al. 1993), and in nor<strong>the</strong>rn Arizona<br />

(Olson et al. 1993) and in <strong>the</strong> Tularosa Mountains<br />

of west-central New Mexico in 1991<br />

(Seamans et al. 1993).<br />

To document <strong>the</strong> current (1990-1993)<br />

distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl, we<br />

defined an owl site as a visual sighting of at least<br />

one adult spotted owl or as a minimum of two<br />

auditory detections in <strong>the</strong> same vicinity in <strong>the</strong><br />

same year. Observations prior to 1990 are<br />

considered historical records <strong>for</strong> <strong>the</strong> purposes of<br />

this report. The methods and limitations of <strong>the</strong>se<br />

data are discussed fur<strong>the</strong>r in <strong>the</strong> White et al.<br />

1995.<br />

HISTORICAL HISTORICAL DISTRIBUTION<br />

DISTRIBUTION<br />

We compiled 600 and 35 historical records<br />

of <strong>Mexican</strong> spotted owls in <strong>the</strong> United States<br />

and Mexico, respectively (Table 1.1). We refer to<br />

<strong>the</strong>se as records and not as independent sites<br />

because several observations may have been<br />

tallied <strong>for</strong> <strong>the</strong> same site. Incomplete in<strong>for</strong>mation<br />

of <strong>the</strong> owls’ locations prevented us from assigning<br />

each record to an individual site. Thus, <strong>the</strong>se<br />

records cannot be used to estimate historical<br />

abundance and are presented only to show <strong>the</strong><br />

approximate extent of <strong>the</strong> owl’s distribution<br />

be<strong>for</strong>e 1990.


Volume II/Chapter 1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 1.1. 1.1. Historical records and minimum numbers of <strong>Mexican</strong> spotted owls found during planned<br />

surveys, and incidental observations by <strong>Recovery</strong> Unit and land ownership.<br />

<strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

UNITED UNITED STATES<br />

STATES<br />

Number Number of of owl owl records records<br />

records<br />

be<strong>for</strong>e be<strong>for</strong>e 1990 1990a<br />

Number Number of of owl owl sites sites<br />

sites<br />

1990 1990 - - - 1993<br />

1993<br />

FS 21 16<br />

BLM 6 10<br />

NPS 34 23<br />

Tribal 20 13b New Mexico State 1 0<br />

Unknownc 5 0<br />

Subtotal Subtotal<br />

87 87<br />

62<br />

62<br />

FS 2 8<br />

BLM 0 6<br />

NPS 0 0<br />

Tribal 1 --b Unknownc 17 0<br />

Subtotal Subtotal<br />

20 20<br />

14<br />

14<br />

FS 25 34<br />

NPS 3 0<br />

New Mexico State 1 0<br />

Private 4 0<br />

Unknownc Colorado Plateau<br />

Sou<strong>the</strong>rn Rocky Mountains - Colorado<br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico<br />

8 0<br />

Subtotal Subtotal<br />

41 41<br />

34<br />

34<br />

Upper Gila Mountains<br />

FS 138 424<br />

BLM 5 0<br />

NPS 5 0<br />

Tribal 20 --b Private 1 0<br />

Unknownc 104 0<br />

Subtotal Subtotal<br />

253 253<br />

424<br />

424<br />

FS 82 97<br />

NPS 13 0<br />

Tribal 0 --b DOD 9 6<br />

Private 8 0<br />

Unknownc Basin and Range - West<br />

57 0<br />

Subtotal Subtotal<br />

169 169<br />

103<br />

103<br />

2


Table Table 1.1. 1.1. 1.1. (continued)<br />

<strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

Volume II/Chapter 1<br />

Number Number of of of owl owl records<br />

records<br />

be<strong>for</strong>e be<strong>for</strong>e 1990 1990 1990a<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Number Number of of owl owl sites sites<br />

sites<br />

1990 1990 - - 1993<br />

1993<br />

FS 18 111<br />

BLM 1 0<br />

NPS 6 10<br />

Tribal 2 --b UNITED UNITED STATES, STATES, continued continued<br />

continued<br />

Basin and Range - East<br />

FWS 1 0<br />

Private 2 0<br />

Subtotal Subtotal<br />

30 30<br />

121<br />

121<br />

United United States States Total<br />

Total<br />

600 600<br />

758<br />

758<br />

Coahuila 4 0<br />

Nuevo Leon 4 1<br />

Tamaulipas 0 0d MEXICO<br />

MEXICO<br />

Sonora 8 9<br />

Chihuahua 10 8<br />

Subtotal Subtotal Subtotal<br />

8 1<br />

d<br />

Sierra Madre Occidental - Norte<br />

Sinaloa 1 0<br />

Subtotal Subtotal<br />

Sierra Madre Oriental - Norte<br />

19 19<br />

17<br />

17<br />

Coahuila<br />

Sierra Madre Occidental - Sur<br />

2 0<br />

Durango 2 0<br />

Aguascalientes 0 1d Zacatecas 0 0d San Luis Potosi 1 0<br />

Guanajuato 1 0<br />

Subtotal Subtotal<br />

Sierra Madre Oriental - Sur<br />

4 1<br />

Eje Neovolcanico<br />

Jalisco 1 0<br />

Colima 1e 0<br />

Michoacan 1 0<br />

Puebla 1e 0<br />

Subtotal Subtotal<br />

2 0<br />

Mexico Mexico Total<br />

Total<br />

35 35<br />

19<br />

19<br />

a Values do not connote numbers of owls nor owl sites because multiple records may exist from <strong>the</strong> same site through time.<br />

b Additional owls are known to exist on many Tribal lands but <strong>the</strong> exact number is unavailable.<br />

c Locations of <strong>the</strong>se records were insufficient <strong>for</strong> assigning a land ownership.<br />

d Additional sightings have been reported from 1994 surveys.<br />

e Unverified record not included in totals (see text).<br />

3


The general vicinity of all historical records<br />

is presented in <strong>the</strong> following section according to<br />

RU and land ownership. This in<strong>for</strong>mation<br />

should help land managers to identify areas that<br />

may be occupied by <strong>Mexican</strong> spotted owls.<br />

Historical records <strong>for</strong> owls in <strong>the</strong> United States<br />

were compiled from several sources which are<br />

cited below. In contrast, much of <strong>the</strong> historical<br />

in<strong>for</strong>mation <strong>for</strong> Mexico was taken from a comprehensive<br />

summary by Williams and Skaggs<br />

(1993).<br />

Colorado Colorado Plateau Plateau <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

Prior to 1990, <strong>Mexican</strong> spotted owls were<br />

recorded from Zion (Kertell 1977, Rinkevich<br />

1991), Canyonlands, Capitol Reef (Sue Linner,<br />

FWS, Salt Lake City Utah, pers. comm.), Mesa<br />

Verde (Reynolds and Johnson 1994), and Grand<br />

Canyon National Parks (McDonald et al. 1991),<br />

Glen Canyon National Recreation Area (Behle<br />

1960), and <strong>the</strong> Cedar City, Richfield, and Vernal<br />

Districts of <strong>the</strong> BLM (Behle 1981; Sue Linner,<br />

FWS, Salt Lake City, Utah, pers. comm.). Table<br />

1.1 presents more detail regarding <strong>the</strong>se records.<br />

The physical attributes of <strong>the</strong>se historical owl<br />

sites include steep-sided, narrow-walled, or<br />

hanging canyons. The biotic attributes include<br />

coniferous overstory in canyon bottoms with an<br />

understory comprised by Gambel oak, bigtooth<br />

maple, boxelder, and scattered aspen groves near<br />

water (McDonald et al. 1991).<br />

Historical accounts also place <strong>the</strong> <strong>Mexican</strong><br />

spotted owl on <strong>the</strong> Kaibab Plateau in Arizona<br />

(Ganey and Balda 1989, North Kaibab Ranger<br />

District, unpublished data), plus many sites in<br />

New Mexico. These sites include Fence Lake<br />

(State), Frances Canyon (BLM), <strong>the</strong> Zuni<br />

Mountains and Mount Taylor (Cibola NF), and<br />

within <strong>the</strong> Zuni and Navajo Nations (McDonald<br />

et al. 1991, New Mexico Natural Heritage Data<br />

Base, Nature Conservancy, Albuquerque, NM).<br />

Generally, vegetation types reported <strong>for</strong> <strong>the</strong>se<br />

areas include montane coniferous <strong>for</strong>ests within<br />

canyon settings. The owl has been observed in<br />

steep-walled canyons with minimal vegetation as<br />

well as in <strong>for</strong>ested, steep-sloped canyons on<br />

Black Mesa and in <strong>the</strong> Chuska Mountains.<br />

Volume II/Chapter 1<br />

4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Sou<strong>the</strong>rn Sou<strong>the</strong>rn Rocky Rocky Mountains Mountains -<br />

-<br />

Colorado Colorado <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

Eighteen historical records of spotted owls<br />

exist within this unit (Webb 1983, Reynolds<br />

1989). Most of <strong>the</strong>se owls were found along <strong>the</strong><br />

Colorado Front Range extending northward to<br />

Fort Collins. Two additional observations, one<br />

each from Rio Grande and San Juan National<br />

Forests, plus one from <strong>the</strong> Sou<strong>the</strong>rn Ute Reservation<br />

were recorded during 1989 surveys<br />

(Reynolds and Johnson 1994; Table 1.1).<br />

Historical owl locations in this recovery unit<br />

occurred in steep-sided canyons. These canyons<br />

are typically broader with walls that are not as<br />

vertical as sites occupied by owls in sou<strong>the</strong>rn<br />

Utah (Colorado Plateau RU). Nor<strong>the</strong>rn aspects<br />

of <strong>the</strong>se canyons contain mixed-conifer <strong>for</strong>est,<br />

while sou<strong>the</strong>rn aspects contain ponderosa pine<br />

and pinyon-juniper. Canyon bottoms contain<br />

Gambel oak and boxelder. <strong>Owl</strong> sightings in<br />

southwestern Colorado were generally in canyons<br />

that cut into mesas covered with pinyon<br />

and juniper. These canyon bottoms contained<br />

mixed-conifer or ponderosa pine-Gambel oak<br />

<strong>for</strong>ests (McDonald et al. 1991).<br />

Sou<strong>the</strong>rn Sou<strong>the</strong>rn Sou<strong>the</strong>rn Rocky Rocky Mountains Mountains -<br />

-<br />

New New Mexico Mexico <strong>Recovery</strong> <strong>Recovery</strong> <strong>Recovery</strong> Unit Unit<br />

Unit<br />

<strong>Mexican</strong> spotted owls are known historically<br />

(Table 1.1) from private and National Forest<br />

lands in <strong>the</strong> San Juan, Sangre de Cristo, and<br />

Jemez Mountains, and near Taos and Sante Fe,<br />

New Mexico (Johnson and Johnson 1985,<br />

McDonald et al. 1991). Incidental observations<br />

between 1979 and 1984 established <strong>the</strong> presence<br />

of <strong>the</strong> owl at nine additional sites in <strong>the</strong> Jemez<br />

Mountains, Sante Fe National Forest (Johnson<br />

and Johnson 1985). The owl has also been<br />

observed in Bandelier National Monument and<br />

near Morhy Lake on State lands (Johnson and<br />

Johnson 1985). Historical spotted owl locations<br />

throughout all of New Mexico (including <strong>the</strong><br />

Basin and Range - East RU) have been described<br />

as “deep, narrow, timbered canyons with cool<br />

shady places, at elevations ranging from 6,500<br />

[1,982 sic] to 9,000 ft [2,744 m],” (McDonald<br />

et al. 1991, after Ligon 1926).


Upper Upper Gila Gila Mountains Mountains <strong>Recovery</strong> <strong>Recovery</strong> Unit Unit<br />

Unit<br />

Prior to a statewide survey conducted from<br />

1984 through 1988, only one <strong>Mexican</strong> spotted<br />

owl was recorded from <strong>the</strong> nor<strong>the</strong>rn Arizona<br />

(San Francisco Peaks) portion of this RU (Huey<br />

1930). In contrast, several owls were reported <strong>for</strong><br />

<strong>the</strong> National Forest lands in <strong>the</strong> Mogollon<br />

Highlands of east-central Arizona and westcentral<br />

New Mexico (Ligon 1926, Skaggs 1988).<br />

Forests in those reports include <strong>the</strong> Apache-<br />

Sitgreaves, Gila, and Cibola National Forests<br />

(Table 1.1). A few observations were also recorded<br />

on BLM land near Bitter Creek, Grant<br />

County and at <strong>the</strong> Gila Cliff Dwellings National<br />

Monument (NPS), both in New Mexico. Following<br />

<strong>the</strong>ir survey of Arizona, Ganey and Balda<br />

(1989) reported 69 sites in <strong>the</strong> Arizona portion<br />

of <strong>the</strong> Mogollon Rim and in nor<strong>the</strong>rn Arizona,<br />

including <strong>the</strong> Coconino, Kaibab, Tonto, and<br />

Apache-Sitgreaves National Forests. Ano<strong>the</strong>r 44<br />

records exist in <strong>the</strong> Arizona Heritage Data Base<br />

(Arizona Game and Fish Department, Phoenix,<br />

AZ). However, <strong>the</strong> latter records are distributed<br />

among habitats similar to those reported by<br />

Ganey and Balda (1989) and include several of<br />

<strong>the</strong> same sites.<br />

Site characteristics <strong>for</strong> <strong>the</strong> historical Arizona<br />

locations (Ganey and Balda 1989) reflect <strong>the</strong><br />

features described <strong>for</strong> o<strong>the</strong>r RU’s: mountain<br />

slopes with mixed-coniferous <strong>for</strong>est, steep-walled<br />

canyons, or ponderosa pine-Gambel oak <strong>for</strong>est at<br />

elevations ranging from 1,525 to 2,925 m<br />

(5,000 to 9,590 ft). In sou<strong>the</strong>rn New Mexico,<br />

Skaggs (1988) found cliffs present at 15 of <strong>the</strong><br />

18 historical sites which he examined. However,<br />

it is unclear how many of <strong>the</strong>se sites were located<br />

in <strong>the</strong> Upper Gila Mountains RU. Skaggs (1988)<br />

also noted a well developed understory of<br />

bigtooth maple and Gambel oak dominated by<br />

mixed-conifer.<br />

Basin Basin and and Range Range - - West West <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

Historical records <strong>for</strong> <strong>Mexican</strong> spotted owls<br />

in this RU include observations in <strong>the</strong> Huachuca<br />

and Chiricahua Mountains during <strong>the</strong> 1890s<br />

(reviewed in McDonald et al. 1991). These birds<br />

were observed in a foothills-oak woodland and<br />

Volume II/Chapter 1<br />

5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

in a fir tree in Pinery Canyon, respectively. Two<br />

o<strong>the</strong>r sightings were recorded in lowland riparian<br />

communities including an owl nesting in cottonwoods<br />

northwest of Tucson in 1872 and near <strong>the</strong><br />

Salt River in 1910 (Bendire 1892, Phillips et al.<br />

1964, McDonald et al. 1991).<br />

More recent surveys found <strong>the</strong> owl occurring<br />

at 84 sites throughout sou<strong>the</strong>rn Arizona (Ganey<br />

and Balda 1989). <strong>Owl</strong>s were located in rocky<br />

canyons or in several <strong>for</strong>est types at elevations<br />

ranging from 1,125 to 2,930 m (3,690 to 9,610<br />

ft) in <strong>the</strong> Atascosa-Pajarito, Santa Rita, Santa<br />

Catalina, Patagonia, Whetstone, Galiuro,<br />

Huachuca, Chiricahua, Pinaleno, Superstition,<br />

Sierra Ancha, Mazatzal, and Bradshaw Mountains,<br />

Arizona. Below 1,300 m (4,264 ft), spotted<br />

owls were found in steep canyons containing<br />

cliffs and stands of live oak, <strong>Mexican</strong> pine and<br />

broad-leaved riparian vegetation (Ganey and<br />

Balda 1989). Above 1,800 m (5,904 ft) owls<br />

were found in mixed-conifer and pine-oak<br />

<strong>for</strong>ests. Mid-elevation observations included sites<br />

with Arizona cypress and <strong>the</strong> o<strong>the</strong>r <strong>for</strong>est types<br />

previously mentioned. The Arizona Heritage<br />

Data Base reports 78 additional records in many<br />

of <strong>the</strong> same mountain ranges from 1974 to<br />

1989. Historical records on private land include<br />

observations near Animas Peak, Black Bill<br />

Spring, and at <strong>the</strong> Gray Ranch, in New Mexico<br />

(Skaggs 1988).<br />

Basin Basin and and Range Range - - East East East <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

Historical locations of <strong>Mexican</strong> spotted owls<br />

occur on lands of several jurisdictions in New<br />

Mexico: <strong>the</strong> Organ Mountains and near Bitter<br />

Creek (BLM); <strong>the</strong> Sandia, Manzano, Sacramento,<br />

and Guadalupe Mountains in <strong>the</strong> Cibola<br />

and Lincoln National Forests; and Carlsbad<br />

National Park (Skaggs 1988). The owl has also<br />

been found in Guadalupe National Park and on<br />

private land in <strong>the</strong> Davis Mountains of Texas<br />

(McDonald et al. 1991, Steve Runnels, The<br />

Heard Natural Science Museum and Wildlife<br />

Sanctuary, McKinney, TX, pers. comm.). One<br />

observation each was also reported on <strong>the</strong> lands<br />

of <strong>the</strong> Mescalero Apache and at <strong>the</strong> Santo<br />

Domingo Pueblo (New Mexico Natural Heritage<br />

Data Base, Nature Conservancy, Albuquerque,


NM). Physical and biotic characteristics were<br />

not documented <strong>for</strong> <strong>the</strong>se various sites. However,<br />

minimal notations describe mixed-conifer <strong>for</strong>ests<br />

on eastern slopes of <strong>the</strong> Sacramento Mountains<br />

(Skaggs 1988). Canyons were mentioned <strong>for</strong><br />

sites near <strong>the</strong> New Mexico-Texas border of <strong>the</strong><br />

Guadalupe Mountains (McDonald et al. 1991).<br />

Sierra Sierra Madre Madre Occidental Occidental -<br />

-<br />

Norte Norte <strong>Recovery</strong> <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

More than half of all historical records of<br />

<strong>Mexican</strong> spotted owls occurring in Mexico have<br />

been reported in this <strong>Recovery</strong> Unit (Table 1.1).<br />

<strong>Owl</strong>s have been recorded from eight locations in<br />

<strong>the</strong> State of Sonora prior to 1990. These birds<br />

occurred in <strong>the</strong> Sierras Pinitos, Azul, de los Ajos,<br />

San Luis, Aconchi, Oposura, and Huachinera<br />

(Williams and Skaggs 1993). The eighth record<br />

was reported from a ridge north of La Mesa,<br />

Mexico. All of <strong>the</strong>se areas are physiographically<br />

and biotically similar to <strong>the</strong> Sky Island Mountains<br />

of sou<strong>the</strong>astern Arizona (Cirett and Diaz<br />

1993). General elevations at or near <strong>the</strong>se<br />

sightings range from 1,950-2,340 m (6,500 to<br />

7,800 ft). Descriptions of sites at or near <strong>the</strong><br />

recorded owls vary. Some descriptions include<br />

dense oaks, pine <strong>for</strong>est, pine-oak woodland,<br />

cliffs, and a spring with alders and sycamores<br />

(Williams and Skaggs 1993).<br />

Ten historical records have been reported<br />

from <strong>the</strong> State of Chihuahua (Table 1.1). <strong>Owl</strong>s<br />

have been collected, observed, or heard near<br />

Sierra Carcay, Arroyo Tinaja, Pacheco, Sierra<br />

Azul, Colonia Garcia, Sierra del Nido, Rancho<br />

La Estancia, Yaguirachic, Pinos Altos, and<br />

Vosagota (Williams and Skaggs 1993). Elevations<br />

at or near <strong>the</strong>se sightings range from<br />

1,710-2,700 m (5,700 to 9,000 ft). Most of<br />

<strong>the</strong>se observations were made in or near pineoak<br />

woodlands (Williams and Skaggs 1993).<br />

<strong>Owl</strong>s have also been reported in canyons with<br />

oaks and madrone, and in subalpine <strong>for</strong>est<br />

comprised of Douglas-fir, true firs, oak, alder,<br />

pines, and chokecherry.<br />

In Sinaloa, one spotted owl was observed<br />

near Rancho Liebre Barranca (Williams and<br />

Skaggs 1993). This area consists of deeply<br />

dissected barrancas with pine and oak <strong>for</strong>est at<br />

Volume II/Chapter 1<br />

6<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

higher altitudes and a mixture of temperate and<br />

tropical <strong>for</strong>est in canyon bottoms.<br />

Sierra Sierra Madre Madre Oriental Oriental -<br />

-<br />

Norte Norte <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

<strong>Spotted</strong> owls have only been reported from<br />

two sites within this <strong>Recovery</strong> Unit. Both<br />

records are from <strong>the</strong> Sierra la Madera of central<br />

Coahuila (Table 1.1). One owl was observed<br />

roosting in a “cliff-lined canyon bottom under a<br />

dense canopy of maples and oaks” in Canada el<br />

Agua (Williams and Skaggs 1993). Ano<strong>the</strong>r owl<br />

was observed and heard in a “garden-like” arroyo<br />

containing pines, oaks, and madrones (Williams<br />

and Skaggs 1993). Elevations of <strong>the</strong>se observations<br />

were approximately 1,900 and 2,100 m<br />

(6,200 and 7,000 ft), respectively.<br />

Sierra Sierra Madre Madre Occidental Occidental Occidental -<br />

-<br />

Sur Sur <strong>Recovery</strong> <strong>Recovery</strong> Unit Unit<br />

Unit<br />

Four historical records exist from <strong>the</strong> States<br />

of Durango, San Luis Potosi, and Guanajuato<br />

(Table 1.1). In Durango, spotted owls have been<br />

observed near Espinazo del Diablo in mixedconifer<br />

<strong>for</strong>est and on two occasions in <strong>the</strong><br />

Michilia Biosphere Reserve. One of <strong>the</strong> owls<br />

found in <strong>the</strong> reserve was observed roosting in a<br />

large oak that was in a cool, wet ravine. The<br />

second owl was found in a pine-oak <strong>for</strong>est.<br />

Remaining records are of two spotted owls<br />

collected at approximately 2,400 m (8,000 ft)<br />

near Cerro Campanario, San Luis Potosi, and<br />

ano<strong>the</strong>r that was collected in <strong>the</strong> State of<br />

Guanajuato (Williams and Skaggs 1993).<br />

Sierra Sierra Madre Madre Oriental Oriental Oriental -<br />

-<br />

Sur Sur <strong>Recovery</strong> <strong>Recovery</strong> Unit<br />

Unit<br />

In this <strong>Recovery</strong> Unit, eight records of<br />

spotted owls have been reported from two States,<br />

Coahuila and Nuevo Leon (Table 1.1). <strong>Owl</strong>s<br />

have been heard and observed on several occasions<br />

east of Saltillo, Coahuila. Elevations of<br />

<strong>the</strong>se records are generally higher than o<strong>the</strong>r<br />

historical sightings and range from 2,700-3,060<br />

m (9,000 to 10,200 ft). The vegetation at <strong>the</strong>se<br />

sites has been described as oak-pine-conifer


woodland or as oak-pine woodland. Cliffs are<br />

present at all sites and chaparral or desert scrub<br />

vegetation can also be seen at two of <strong>the</strong>se sites<br />

(Williams and Skaggs 1993). In addition, two of<br />

<strong>the</strong> four historical records from Nuevo Leon are<br />

from <strong>the</strong> mountains east of Saltillo and south of<br />

Monterrey. At one of <strong>the</strong>se sites a spotted owl<br />

was heard calling from an oak-pine-conifer<br />

<strong>for</strong>est, approximately 2,550 m (8,500 ft) in<br />

elevation. The second was heard calling from an<br />

“oak-pine woodland mixed with Tamaulipan<br />

thorn woodland and scrub and palms on <strong>the</strong><br />

canyon cliffs” (Williams and Skaggs 1993). The<br />

elevation at this site was 1,350 m (4,500 ft), <strong>the</strong><br />

lowest to be recorded <strong>for</strong> spotted owls in<br />

Mexico. The remaining two Nuevo Leon records<br />

are from <strong>the</strong> slopes of Cerro Potosi (Williams<br />

and Skaggs 1993). One female spotted owl was<br />

collected at 2,250 m (7,500 ft) in 1946. Ano<strong>the</strong>r<br />

owl was heard in 1978 from a pine woodland<br />

near <strong>the</strong> summit at 3,630 m (12,100 ft), <strong>the</strong><br />

highest elevation recorded <strong>for</strong> a <strong>Mexican</strong> spotted<br />

owl.<br />

Eje Eje Neovolcanico Neovolcanico <strong>Recovery</strong> <strong>Recovery</strong> Unit Unit<br />

Unit<br />

Only two confirmed records of <strong>Mexican</strong><br />

spotted owls exist <strong>for</strong> <strong>the</strong> sou<strong>the</strong>rnmost <strong>Recovery</strong><br />

Unit (Table 1.1). Specimens of spotted owls have<br />

been collected from <strong>the</strong> States of Jalisco and<br />

Michoacan. Two o<strong>the</strong>r records from <strong>the</strong> States of<br />

Colima and Puebla have been published<br />

(Enriquez-Rocha et al. 1993) but have not been<br />

verified (Williams and Skaggs 1993).<br />

In Jalisco, <strong>the</strong> owls were found on <strong>the</strong> north<br />

slope of Cerro Nevado de Colima in a park-like<br />

pine <strong>for</strong>est with broad-leaved oaks, and mesic<br />

ground flora near 2,400 m (8,000 ft) elevation<br />

(Williams and Skaggs 1993). In Michoacan, <strong>the</strong><br />

holotype of <strong>the</strong> subspecies and only existing<br />

State record was collected in 1903 from Cerro<br />

Tancitaro above 1,950 m (6,500 ft). Details<br />

regarding <strong>the</strong> spotted owls allegedly collected in<br />

Colima and Puebla have not been reported<br />

(Enriquez-Rocha et al. 1993). We mention <strong>the</strong>se<br />

latter records because <strong>the</strong>y suggest possible<br />

extensions of <strong>the</strong> owl’s range. However, we have<br />

not included <strong>the</strong>m in <strong>the</strong> totals presented in<br />

Table 1.1.<br />

Volume II/Chapter 1<br />

7<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

CURRENT CURRENT DISTRIBUTION<br />

DISTRIBUTION<br />

AND AND ABUNDANCE<br />

ABUNDANCE<br />

Number Number of of Sites<br />

Sites<br />

Surveys <strong>for</strong> <strong>Mexican</strong> spotted owls conducted<br />

from 1990 through 1993 indicate that <strong>the</strong><br />

species persists in most locations reported prior<br />

to 1989. Notable exceptions include riparian<br />

habitats in <strong>the</strong> lowlands of Arizona and New<br />

Mexico, and all previously occupied areas in <strong>the</strong><br />

sou<strong>the</strong>rn States of Mexico. As a result of planned<br />

surveys, additional sightings have been reported<br />

<strong>for</strong> all recovery units. New locations will undoubtedly<br />

be reported following future surveys.<br />

The current known range of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl extends north from Aguascalientes,<br />

Mexico, through <strong>the</strong> mountains of Arizona, New<br />

Mexico, and western Texas to <strong>the</strong> canyons of<br />

sou<strong>the</strong>rn Utah, southwestern Colorado, and <strong>the</strong><br />

Front Range of central Colorado (Figures 1.1<br />

and 1.2). Results from planned surveys and<br />

incidental observations conducted during 1990<br />

through 1993 indicate one or more owls have<br />

been observed at a minimum of 758 sites in <strong>the</strong><br />

United States and 19 sites in Mexico (Table 1.1).<br />

The greatest concentration of <strong>the</strong> known<br />

sites in <strong>the</strong> United States occurs in <strong>the</strong> Upper<br />

Gila Mountains <strong>Recovery</strong> Unit (55.9%) followed<br />

by <strong>the</strong> Basin and Range-East (16.0%),<br />

and Basin and Range-West (13.6%), Colorado<br />

Plateau (8.2%), Sou<strong>the</strong>rn Rocky Mountains -<br />

New Mexico (4.5%), and Sou<strong>the</strong>rn Rocky<br />

Mountains - Colorado (1.8%) <strong>Recovery</strong> Units.<br />

Thus, fewer owl sites are currently known to<br />

occur north of <strong>the</strong> Upper Gila Mountains<br />

<strong>Recovery</strong> Unit (12.7%) than to <strong>the</strong> south of this<br />

recovery unit (29.6%). In Mexico, <strong>the</strong> majority<br />

of spotted owls have been documented in <strong>the</strong><br />

Sierra Madre Occidental - Norte RU (89.5%).<br />

However, <strong>the</strong> number of identified sites within a<br />

given RU depends on survey intensity <strong>for</strong> which<br />

we have little reliable data. There<strong>for</strong>e, <strong>the</strong> percentages<br />

of sites within a given RU may not<br />

reflect <strong>the</strong> true relative abundance of <strong>Mexican</strong><br />

spotted owls.


Volume II/Chapter 1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 1.1. 1.1. <strong>Recovery</strong> unit boundaries and current distribution of <strong>Mexican</strong> spotted owls in <strong>the</strong><br />

United States based on planned surveys and incidental observations recorded from 1990 through 1993.<br />

FPO<br />

8


Volume II/Chapter 1<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 1.2. 1.2. Current distribution of <strong>Mexican</strong> spotted owls in Mexico based on planned surveys and<br />

incidental observations recorded from 1990 through 1993.<br />

9


Number Number of of <strong>Owl</strong>s<br />

<strong>Owl</strong>s<br />

A reliable estimate of <strong>the</strong> number of <strong>Mexican</strong><br />

spotted owls throughout its entire range is not<br />

available. Fletcher (1990) calculated that 2,074<br />

owls existed in Arizona and New Mexico in<br />

1990 using in<strong>for</strong>mation ga<strong>the</strong>red by <strong>the</strong> FS,<br />

Southwestern Region. McDonald et al. (1991)<br />

modified Fletcher’s (1990) calculations reporting<br />

a total of 2,160 owls in <strong>the</strong> United States. If one<br />

assumes that all 758 sites included in this recovery<br />

plan were occupied by owl pairs, <strong>the</strong>n at least<br />

1,516 adult or subadult owls were known to<br />

exist in <strong>the</strong> United States and 38 adult or subadult<br />

owls in Mexico from 1990 through 1993.<br />

These numbers are not reliable estimates of<br />

current population size because no measures of<br />

bias or precision can be produced. Fur<strong>the</strong>r, <strong>the</strong><br />

amount of survey ef<strong>for</strong>t devoted to deriving<br />

<strong>the</strong>se numbers cannot be reliably calculated, nor<br />

is an accurate measure available <strong>for</strong> areas or<br />

habitats surveyed. Thus, we did not believe it<br />

would be useful to estimate <strong>the</strong> size of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl population given <strong>the</strong><br />

limited quality of data currently available. At<br />

best, our total numbers reported in Table 1.1<br />

represent a range <strong>for</strong> <strong>the</strong> minimum number of<br />

owls known to exist during some portion of a<br />

four year period in <strong>the</strong> United States and Mexico<br />

(777 individuals if each site was occupied by a<br />

single owl to 1,554 individuals if each site was<br />

occupied by a pair).<br />

Density<br />

Density<br />

The abundance of any terrestrial organism is<br />

more appropriately presented as density, <strong>the</strong><br />

number of individuals per unit of area (Caughley<br />

1977), hereafter referred to as “crude density”<br />

(Franklin et al. 1990). Because <strong>the</strong> <strong>Mexican</strong><br />

spotted owl occupies a variety of habitats<br />

throughout its range, ecological density, <strong>the</strong><br />

number of individuals per area of usable habitat,<br />

(Tanner 1978) would be a more meaningful<br />

measure of abundance. At this time, rangewide<br />

estimates of ei<strong>the</strong>r crude or ecological density of<br />

<strong>Mexican</strong> spotted owls cannot be provided <strong>for</strong> <strong>the</strong><br />

same reasons that population numbers cannot be<br />

estimated reliably.<br />

Volume II/Chapter 1<br />

10<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

^<br />

Two estimates of crude density (D) exist<br />

from studies conducted at ei<strong>the</strong>r end of <strong>the</strong><br />

Upper Gila Mountains <strong>Recovery</strong> Unit. These<br />

estimates were reported <strong>for</strong> <strong>the</strong> Coconino Study<br />

Area (CSA) in nor<strong>the</strong>rn Arizona and <strong>for</strong> <strong>the</strong> Gila<br />

Study Area (GSA) in west-central New Mexico<br />

by Gutiérrez et al. (1994). Density of adult and<br />

subadult owls in both studies was estimated<br />

using a count of individuals divided by <strong>the</strong> size<br />

of <strong>the</strong> study area (CSA: 484 km2 [187 mi2 ];<br />

GSA: 323 km2 [125 mi2 ]). Identity of owls was<br />

established by capturing and marking or by<br />

direct observation at daytime roosts when<br />

marking was not possible. Methods were similar<br />

to those reported <strong>for</strong> nor<strong>the</strong>rn spotted owls by<br />

Franklin et al. (1990) and Ward et al. (1991) and<br />

resulted only in density of territorial individuals.<br />

Only <strong>the</strong> 1993 estimates of density (D) are<br />

presented here because <strong>the</strong> boundaries of <strong>the</strong><br />

CSA study area were shifted between 1991 and<br />

1992.<br />

D of <strong>Mexican</strong> spotted owls in 1993 was<br />

0.120 owls/km2 (0.310 owls/mi2 ) in <strong>the</strong> CSA<br />

and 0.180 owls/km2 (0.464 owls/mi2 ) in <strong>the</strong><br />

GSA. The CSA has more ponderosa pine-<br />

Gambel oak <strong>for</strong>est (72.6%) and less mixedconifer<br />

<strong>for</strong>est (14.4%) than <strong>the</strong> GSA (22.3%<br />

and 28.5%, respectively; Gutiérrez et al. 1994).<br />

The larger proportion of mixed-conifer may<br />

partially explain higher owl density in <strong>the</strong> GSA.<br />

For comparison (Figure 1.3a), 1993 density of<br />

Cali<strong>for</strong>nia spotted owls in <strong>the</strong> San Bernardino<br />

Mountains, Cali<strong>for</strong>nia (LaHaye and Gutiérrez<br />

1994) was 0.118 owls/km2 (0.305 owls/mi2 ) and<br />

density of nor<strong>the</strong>rn spotted owls in northwestern<br />

Cali<strong>for</strong>nia (Franklin, unpublished data) was<br />

0.272 + 0.004 (SE) owls/km2 (0.703 + 0.009<br />

owls/mi2 ). Survey methods used in <strong>the</strong>se later<br />

two studies were identical to those used in <strong>the</strong><br />

CSA and GSA. Naive density estimates, <strong>the</strong><br />

number of owls counted divided by study area<br />

size, were used in this comparison (Figure 1.3a)<br />

except <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn spotted owl population.<br />

In <strong>the</strong> latter case, a Jolly-Seber model was used<br />

to estimate numbers of owls and an associated<br />

sampling variance. The two density estimators<br />

are similar <strong>for</strong> spotted owls (Ward et al. 1991).<br />

Combining <strong>the</strong> CSA and GSA density data<br />

and weighting by area provides an average<br />

estimate of 0.144 owls/km2 (0.372 owls/mi2 ^<br />

^<br />

)


Volume II/Chapter 1<br />

FPO<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure Figure 1.3. 1.3. 1.3. Density of (a) (a) all three subspecies compared among regions and (b) (b) <strong>Mexican</strong> spotted<br />

owls among three <strong>for</strong>est types in <strong>the</strong> Sacramento Mountains, New Mexico, based on Skaggs and Raitt<br />

(1988). Sources of estimates <strong>for</strong> nor<strong>the</strong>rn Arizona (CSA) and west-central New Mexico (GSA) are from<br />

Gutiérrez et al. (1994); sou<strong>the</strong>rn Cali<strong>for</strong>nia are from LaHaye and Gutiérrez (1994); and northwestern<br />

Cali<strong>for</strong>nia are from Franklin (unpublished data). Vertical bars are 95% confidence intervals.<br />

11


within <strong>the</strong> Upper Gila Maintains RU. However,<br />

this estimate should not be extrapolated to a<br />

larger area because (1) <strong>the</strong> CSA and GSA were<br />

not randomly selected and (2) studies were not<br />

sufficiently replicated. Both of <strong>the</strong>se problems<br />

could severely bias extrapolated estimates because<br />

<strong>the</strong> areas studied are not necessarily representative<br />

samples of <strong>the</strong> entire recovery unit or<br />

subspecies’ range.<br />

In ano<strong>the</strong>r study, Skaggs and Raitt (1988)<br />

examined <strong>the</strong> density of <strong>Mexican</strong> spotted owls<br />

among three <strong>for</strong>est types in <strong>the</strong> Sacramento<br />

Mountains (Basin and Range - East RU). Eighteen<br />

23.1- km 2 (9-mi 2 ) quadrats, six in each<br />

<strong>for</strong>est type, were surveyed <strong>for</strong> spotted owls.<br />

Quadrats were classified as pinyon-juniper<br />

woodland, pine, or mixed-conifer <strong>for</strong>est according<br />

to <strong>the</strong> most common tree species within <strong>the</strong><br />

quadrat. <strong>Owl</strong> density averaged across <strong>the</strong> three<br />

<strong>for</strong>est types (x = 0.126 owls/km 2 [0.325<br />

owls/mi 2 ]) was similar to <strong>the</strong> average estimate<br />

from <strong>the</strong> two southwestern demographic studies.<br />

When partitioned by <strong>for</strong>est type, analysis of<br />

<strong>the</strong>se data by <strong>the</strong> Team showed significantly<br />

higher densities (F = 16.93, df = 2, P = 0.0001)<br />

in <strong>the</strong> mixed-conifer (x = 0.275 owls/km 2 ,<br />

SE = 0.046 [0.704 owls/mi 2 , SE = 0.117]) compared<br />

to <strong>the</strong> pine-dominated (x = 0.080<br />

owls/km 2 , SE = 0.028 [0.204 owls/mi 2 ,<br />

SE = 0.073]) and pinyon-juniper habitats<br />

(x = 0.022 owls/km 2 , SE = 0.036 [0.056 owls/<br />

mi 2 , SE = 0.038]). Density was not statistically<br />

different between <strong>the</strong> latter two <strong>for</strong>est types<br />

(Figure 1.3b). Forest type explained 69.3% of<br />

<strong>the</strong> variation in owl density using this ANOVA<br />

model. Skaggs and Raitt (1988) showed similar<br />

results using density of sites ra<strong>the</strong>r than owl<br />

density.<br />

Volume II/Chapter 1<br />

CONCLUSIONS<br />

CONCLUSIONS<br />

CONCLUSIONS<br />

The <strong>Mexican</strong> spotted owl currently occupies<br />

a broad geographic area, but it does not occur<br />

uni<strong>for</strong>mly throughout its range. Instead, <strong>the</strong> owl<br />

occurs in disjunct localities that correspond to<br />

isolated mountain systems and canyons. This<br />

distribution mimics most historical locations,<br />

with a few exceptions. The owl has not been<br />

reported along major riparian corridors in<br />

12<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Arizona and New Mexico where it historically<br />

occurred, nor in historically-documented areas<br />

of sou<strong>the</strong>rn Mexico. Riparian communities and<br />

previously occupied localities in <strong>the</strong> southwestern<br />

United States and sou<strong>the</strong>rn Mexico have<br />

undergone significant habitat alteration since <strong>the</strong><br />

historical sightings (USDI 1994). However, <strong>the</strong><br />

amount of ef<strong>for</strong>t devoted to surveying <strong>the</strong>se areas<br />

is poorly known and future surveys may document<br />

spotted owls. Surveys conducted to relocate<br />

spotted owls have been unsuccessful in<br />

nor<strong>the</strong>rn Colorado near Fort Collins and Boulder,<br />

where records exist from <strong>the</strong> early 1970s and<br />

1980s, and in <strong>the</strong> Book Cliffs of east-central<br />

Utah where owls were recorded in 1958.<br />

The majority of <strong>Mexican</strong> spotted owls<br />

currently known to exist occur in <strong>the</strong> Upper Gila<br />

Mountains <strong>Recovery</strong> Unit. This unit can be<br />

considered a critical nucleus <strong>for</strong> <strong>the</strong> subspecies<br />

because of its central location within <strong>the</strong> owl’s<br />

range and its seemingly high density of owls.<br />

O<strong>the</strong>r areas likely to be important include <strong>the</strong><br />

Sky-islands of sou<strong>the</strong>astern Arizona and <strong>the</strong><br />

Sacramento Mountains, New Mexico (Basin and<br />

Range RUs). Throughout its range, most (91%)<br />

<strong>Mexican</strong> spotted owls occur on public land<br />

administered by <strong>the</strong> FS.<br />

Density estimates of <strong>Mexican</strong> spotted owls<br />

contrasted among <strong>for</strong>est types in <strong>the</strong> Sacramento<br />

Mountains and between two areas in <strong>the</strong> Upper<br />

Gila Mountains RU suggest that mixed-conifer<br />

supports more owls compared to pine-oak, pine,<br />

and pinyon-juniper <strong>for</strong>est types. <strong>Mexican</strong> spotted<br />

owl densities reported from three areas are<br />

similar to those reported <strong>for</strong> Cali<strong>for</strong>nia spotted<br />

owls occurring in <strong>the</strong> San Bernardino Mountains,<br />

Cali<strong>for</strong>nia and slightly less than <strong>the</strong> density<br />

of nor<strong>the</strong>rn spotted owls occurring in<br />

northwestern Cali<strong>for</strong>nia.<br />

Limited in<strong>for</strong>mation inhibits reliable estimation<br />

of <strong>the</strong> absolute number of <strong>Mexican</strong> spotted<br />

owls. However, it is apparent from current<br />

patterns in distribution and habitat use that <strong>the</strong><br />

subspecies is rare relative to o<strong>the</strong>r raptors and is<br />

distributed discontinuously throughout its<br />

range. Species existing under such conditions are<br />

considered vulnerable to extirpation (see<br />

Dawson et al. 1987 <strong>for</strong> discussion relevant to<br />

spotted owls). Although future ef<strong>for</strong>ts will<br />

undoubtedly discover additional owls, <strong>the</strong> extent


and total number of this subspecies in <strong>the</strong><br />

United States will likely not change in magnitude<br />

enough to alter this conclusion. The contrary<br />

is true <strong>for</strong> Mexico where planned surveys<br />

have begun only recently.<br />

Consequently, current strategies developed<br />

to conserve <strong>Mexican</strong> spotted owls will be<br />

limited to basic knowledge about <strong>the</strong> owl’s<br />

distribution. More specific recommendations<br />

will require additional in<strong>for</strong>mation on <strong>the</strong> total<br />

population size, population structure, or interactions<br />

among subpopulations.<br />

Volume II/Chapter 1<br />

LITERATURE LITERATURE CITED<br />

CITED<br />

Behle, W. H. 1960. The birds of sou<strong>the</strong>astern<br />

Utah. Univ. Utah Biol. Surv. 12:1-56.<br />

–—. 1981. The birds of Nor<strong>the</strong>astern Utah.<br />

Occas. Publ. No. 2, Utah Mus. of Nat. Hist.,<br />

Univ. of Utah. Salt Lake City, Utah. 136pp.<br />

Bendire, C. E. 1892. Life histories of North<br />

American birds. U.S. Nat. Mus. Spec. Bull. 1.<br />

Blakesley, J. B., A. B. Franklin, and R. J.<br />

Gutiérrez. 1992. <strong>Spotted</strong> owl roost and nest<br />

site selection in northwestern Cali<strong>for</strong>nia. J.<br />

Wildl. Manage. 56:388-392.<br />

Caughley, G. 1977. Analysis of vertebrate<br />

populations. John Wiley and Sons, London,<br />

England. 234pp.<br />

Cirett-Galan, J. M., and E. R. Diaz. 1993.<br />

Estatus y distribucion del buho manchado<br />

<strong>Mexican</strong>o (<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>) en Sonora,<br />

Mexico. Centro Ecológico de Sonora. 66pp.<br />

Dawson, W. R., J. D. Ligon, J. R. Murphy, J. P.<br />

Myers, D. Simberloff, and J. Verner. 1987.<br />

Report of <strong>the</strong> scientific advisory panel on <strong>the</strong><br />

spotted owl. Condor 89:205-229.<br />

Duncan, R. B., S. M. Speich, and J. D. Taiz.<br />

1993. Banding and blood sampling study of<br />

<strong>Mexican</strong> spotted owls in sou<strong>the</strong>astern Arizona.<br />

Annual report submitted to Coronado Nat.<br />

For., Tucson, Ariz. 21pp.<br />

Enriquez-Rocha, P., J. L. Rangel-Salazar, and D.<br />

W. Holt. 1993. Presence and distribution of<br />

<strong>Mexican</strong> owls: a review. J. Raptor Res.<br />

27:154-160.<br />

Fletcher, K. W. 1990. Habitats used, abundance<br />

and distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl,<br />

<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>, on National Forest<br />

13<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

system lands. U.S. For. Serv., Southwestern<br />

Region, Albuquerque, N.M. 55pp.<br />

Forsman, E. D., E. C. Meslow, and M. J. Strub.<br />

1977. <strong>Spotted</strong> owl abundance in young versus<br />

old-growth <strong>for</strong>ests, Oregon. Wildl. Soc. Bull.<br />

5:43-47.<br />

Forsman, E. D., C. R. Bruce, M. A. Walter, and<br />

E. C. Meslow. 1987. A current assessment of<br />

<strong>the</strong> spotted owl population in Oregon.<br />

Murrelet 68:51-54.<br />

Franklin, A. B., J. P. Ward, R. J. Gutiérrez, and<br />

G. I. Gould, Jr. 1990. Density of nor<strong>the</strong>rn<br />

spotted owls in northwestern Cali<strong>for</strong>nia. J.<br />

Wildl. Manage. 54:1-10.<br />

Ganey, J. L., and R. P. Balda. 1989. Distribution<br />

and habitat use of <strong>Mexican</strong> spotted owls in<br />

Arizona. Condor 91:355-361.<br />

Gutiérrez, R. J., D. R. Olson, and M. E.<br />

Seamans. 1994. Demography of two <strong>Mexican</strong><br />

spotted owl populations in Arizona and New<br />

Mexico. Report submitted to Humboldt State<br />

Univ. Foundation, Arcata, Calif. 29pp.<br />

Huey, L. M. 1930. Notes from <strong>the</strong> vicinity of<br />

San Francisco Mountain, Arizona. Condor<br />

32:128.<br />

Johnson, J. A., and T. H. Johnson. 1985. The<br />

status of <strong>the</strong> spotted owl in nor<strong>the</strong>rn New<br />

Mexico. New Mexico Dept. Game and Fish,<br />

Sante Fe, N.M. 39pp.<br />

Kertell, K. 1977. The spotted owl at Zion<br />

National Park, Utah. West. Birds 8:147-150.<br />

LaHaye, W. S., and R. J. Gutiérrez. 1994. Big<br />

Bear spotted owl study, 1993. Report<br />

submitted to Calif. Dept. of Fish and Game,<br />

Non-game Bird and Mammal Wildl. Manage.<br />

Div., Sacramento, Calif. 12pp.<br />

Ligon, J. S. 1926. Habits of <strong>the</strong> spotted owl<br />

(Syrnium <strong>occidentalis</strong>). Auk 43:421-429.<br />

McDonald, C. B., J. Anderson, J. C. Lewis, R.<br />

Mesta, A. Ratzlaff, T. J. Tibbitts, and S. O.<br />

Williams. 1991. <strong>Mexican</strong> spotted owl (<strong>Strix</strong><br />

<strong>occidentalis</strong> <strong>lucida</strong>) status report. U.S. Fish and<br />

Wildl. Serv., Albuquerque, N.M. 85pp.<br />

Olson, D. R., M. E. Seamans, and R. J.<br />

Gutiérrez. 1993. Demography of two<br />

<strong>Mexican</strong> spotted owl (<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>)<br />

populations in Arizona and New Mexico:<br />

preliminary results, 1992. Humboldt State<br />

Univ., Arcata, Calif.


Phillips, A. R., J. T. Marshall, Jr., and G.<br />

Monson. 1964. The birds of Arizona. Univ. of<br />

Ariz. Press., Tucson, Ariz.<br />

Reynolds, R. T., 1989. Preliminary survey of <strong>the</strong><br />

distribution and habitat of <strong>Mexican</strong> spotted<br />

owls in Colorado. U.S. For. Serv. Rocky Mtn.<br />

For. and Range Exp. Stn., Laramie, Wyo.<br />

19pp.<br />

——., and C. L. Johnson. 1994. Distribution<br />

and ecology of <strong>the</strong> <strong>Mexican</strong> spotted owl in<br />

Colorado: annual report. U.S. For. Serv.<br />

Rocky Mtn. For. and Range Exp. Stn., Ft.<br />

Collins, Colo. 14pp.<br />

Rinkevich, S. E. 1991. Distribution and habitat<br />

characteristics of <strong>Mexican</strong> spotted owls in<br />

Zion National Park, Utah. M.S. Thesis,<br />

Humboldt State Univ., Arcata, Calif. 62pp.<br />

Seamans, M. E., D. R. Olson, and R. J.<br />

Gutiérrez. 1993. Demography of two<br />

<strong>Mexican</strong> spotted owl (<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>)<br />

populations in Arizona and New Mexico:<br />

preliminary results, 1992. Humboldt State<br />

Univ., Arcata, Calif.<br />

Skaggs, R. W., 1988. Status of <strong>the</strong> <strong>Spotted</strong> <strong>Owl</strong><br />

in sou<strong>the</strong>rn New Mexico: 1900-1987. New<br />

Mexico Dept. Game and Fish, Sante Fe, N.M.<br />

——., and R. J. Raitt. 1988. A spotted owl<br />

inventory of <strong>the</strong> Lincoln National Forest,<br />

Sacramento Division, 1988. New Mexico<br />

Dept. Game and Fish, Sante Fe, N.M. 12pp.<br />

Tanner, J. T. 1978. Guide to <strong>the</strong> study of animal<br />

populations. Univ. Tennessee Press, Knoxville,<br />

Tenn. 186pp.<br />

Volume II/Chapter 1<br />

14<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner. 1990. A<br />

conservation strategy <strong>for</strong> <strong>the</strong> spotted owl. U.S.<br />

Gov. Print. Off., Washington, D.C. 427pp.<br />

USDA Forest Service. 1990. Management<br />

guidelines and inventory and monitoring<br />

protocols <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl in <strong>the</strong><br />

Southwest Region. Federal Register<br />

55:27278-27287.<br />

USDI. 1994. The impact of federal programs on<br />

wetlands. Vol II., A report to Congress by <strong>the</strong><br />

Secretary of <strong>the</strong> Interior. U. S. Dept. of<br />

Interior, Washington, D.C.<br />

Ward, J. P., Jr., A. B. Franklin, and R. J.<br />

Gutiérrez. 1991. Using search time and<br />

regression to estimate abundance of territorial<br />

spotted owls. Ecol. Applic. 1:207-214.<br />

Webb, B. 1983. Distribution and nesting requirements<br />

of montane <strong>for</strong>est owls in Colo<br />

rado. Part IV: spotted owl (<strong>Strix</strong> <strong>occidentalis</strong>).<br />

Colo. Field Ornithol. 17:2-8.<br />

White, G.C., A. Franklin, and J.P. Ward, Jr.,<br />

1995. Population biology. Pages ___ - __ in<br />

<strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Vol. . USDI, U.S. Fish and Wildl. Serv.,<br />

Albuquerque, N.M.<br />

Williams, S. O., III, and R. W. Skaggs. 1993.<br />

Distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl (<strong>Strix</strong><br />

<strong>occidentalis</strong> <strong>lucida</strong>) in Mexico. Unpubl. Tech.<br />

Rep., New Mexico Dept. Game and Fish,<br />

Sante Fe, New Mexico. 38pp.


The development of recovery guidelines<br />

and criteria <strong>for</strong> <strong>Mexican</strong> spotted owl populations<br />

requires understanding <strong>the</strong> life history traits and<br />

population processes <strong>for</strong> this species. We examined<br />

<strong>the</strong> characteristics and trends of <strong>Mexican</strong><br />

spotted owl populations at three spatial scales:<br />

range-wide, regional, and local. Range-wide<br />

scales correspond to characteristics and processes<br />

occurring across <strong>the</strong> U.S. geographic range of<br />

<strong>the</strong> subspecies. Regional scales correspond to<br />

recovery units that were delineated according to<br />

broad ecological patterns (see Rinkevich et al.<br />

1995). Local scales roughly correspond to<br />

subpopulations that occur within each recovery<br />

unit. Ideally, <strong>the</strong> same population characteristics<br />

and processes should be examined over different<br />

temporal and spatial scales. However, temporal<br />

comparisons were restricted because historical<br />

in<strong>for</strong>mation on <strong>Mexican</strong> spotted owl populations<br />

does not exist. Thus, we could not compare<br />

historical and current populations to assess <strong>the</strong><br />

impacts of past and present management activities.<br />

In this section, we first review <strong>the</strong> sources<br />

of in<strong>for</strong>mation available <strong>for</strong> making inferences<br />

about <strong>Mexican</strong> spotted owl populations. We<br />

<strong>the</strong>n quantitatively describe <strong>the</strong> life history<br />

characteristics of <strong>the</strong> owl using age- and sexspecific<br />

survival probabilities and fecundity rates.<br />

These characteristics were estimated from data<br />

collected during radio-telemetry studies, banding<br />

studies, and a regional FS monitoring program.<br />

We examined population trends first by estimating<br />

<strong>the</strong> finite rate of population change (l) on a<br />

local scale using our estimates of survival and<br />

fecundity from selected studies and <strong>the</strong>n by<br />

evaluating temporal trends in occupancy rates of<br />

FS management territories (MTs) on a regional<br />

scale. Finally, we attempted to examine relationships<br />

of vegetation type on reproductive output<br />

at a range-wide scale. Through this stepwise<br />

procedure, we evaluated <strong>the</strong> current state of<br />

<strong>Mexican</strong> spotted owl populations.<br />

Volume II/Chapter 2<br />

CHAPTER CHAPTER 2: 2: Population Population Biology<br />

Biology<br />

Gary C. White, Alan B. Franklin,<br />

and James P. Ward, Jr.<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

SOURCES SOURCES OF OF INFORMATION<br />

INFORMATION<br />

Since 1989, organized surveys <strong>for</strong> <strong>Mexican</strong><br />

spotted owls have been conducted by personnel<br />

of <strong>the</strong> FS, BLM, NPS, Tribes, State wildlife<br />

agencies, and by private researchers. Most<br />

surveys followed <strong>the</strong> procedures described <strong>for</strong> <strong>the</strong><br />

FS Region 3 (USDA Forest Service 1990).<br />

Under <strong>the</strong> FS Region 3 system, two types of<br />

surveys, inventory and monitoring, were conducted<br />

<strong>for</strong> different purposes. Inventories were<br />

general surveys used to detect <strong>the</strong> presence of<br />

spotted owls within a defined area. Monitoring<br />

specifically assessed temporal changes in site<br />

occupancy and reproduction by spotted owls in<br />

management territories (MT). Both procedures<br />

required adherence to a standard survey protocol.<br />

In addition, two types of monitoring,<br />

“<strong>for</strong>mal” and “in<strong>for</strong>mal,” were utilized, with<br />

<strong>for</strong>mal monitoring following guidelines of<br />

USDA Forest Service (1990). MTs were monitored<br />

each year from 1989 through 1993 with<br />

<strong>the</strong> <strong>for</strong>mal procedures if <strong>the</strong>y had (1) no previous<br />

management activity, (2) activity<br />

5-20 years prior to monitoring, or (3) recent<br />

activity within 5 years of monitoring. Formal<br />

monitoring resulted in more survey ef<strong>for</strong>t per<br />

MT than <strong>for</strong> in<strong>for</strong>mal monitoring (although<br />

some in<strong>for</strong>mal monitoring sites were visited<br />

more often than required by <strong>for</strong>mal monitoring).<br />

In<strong>for</strong>mal monitoring could be conducted at<br />

any site not <strong>for</strong>mally monitored in any year. The<br />

greatest amount of ef<strong>for</strong>t <strong>for</strong> surveys was devoted<br />

to <strong>for</strong>mal monitoring, followed by in<strong>for</strong>mal<br />

monitoring, and <strong>the</strong>n inventory.<br />

A limitation of <strong>the</strong> FS database was that <strong>the</strong><br />

MTs surveyed were not randomly sampled from<br />

all possible MTs. MTs were added to <strong>the</strong> database<br />

as owls were found. We do not expect<br />

excessive bias in estimated fecundity or persistence<br />

(as defined in Life History Parameters<br />

section) from this nonrandom sample because<br />

<strong>the</strong>se parameters are probably not directly linked<br />

to <strong>the</strong> sample selection procedures. We do expect


significant biases in <strong>the</strong> estimates of density,<br />

abundance, and occupancy rate because including<br />

an MT in <strong>the</strong> sample is directly linked to<br />

<strong>the</strong>se parameters.<br />

In 1990, a study of <strong>Mexican</strong> spotted owl<br />

population dynamics was initiated in <strong>the</strong> Sky<br />

Islands of sou<strong>the</strong>astern Arizona (Duncan et al.<br />

1993). This study complemented inventory and<br />

monitoring ef<strong>for</strong>ts on <strong>the</strong> Coronado National<br />

Forest. In 1991, two o<strong>the</strong>r studies on population<br />

dynamics were initiated: one on <strong>the</strong> Coconino<br />

National Forest in nor<strong>the</strong>rn Arizona (Gutiérrez<br />

et al. 1993, 1994) and <strong>the</strong> o<strong>the</strong>r on <strong>the</strong> Gila<br />

National Forest in west-central New Mexico<br />

(Gutiérrez et al. 1993, 1994). These two studies<br />

were conducted independently from inventory<br />

and monitoring ef<strong>for</strong>ts, and thus used methods<br />

different than <strong>the</strong> <strong>for</strong>mal monitoring protocol.<br />

In all three population studies, owls were located,<br />

captured, individually marked <strong>for</strong> future<br />

recognition, and <strong>the</strong> number of fledged young<br />

were recorded. These population studies offer<br />

<strong>the</strong> most reliable in<strong>for</strong>mation currently available<br />

<strong>for</strong> estimating abundance, rates of reproduction<br />

and survival, and <strong>for</strong> deriving short-term estimates<br />

of population change. However, inferences<br />

from <strong>the</strong>se studies are somewhat limited because<br />

<strong>the</strong>ir study areas were not randomly selected<br />

from a defined sampling frame. Even with this<br />

limitation, estimates from <strong>the</strong>se studies are useful<br />

in our preliminary evaluation of <strong>Mexican</strong> spotted<br />

owl population biology.<br />

Most inventory and monitoring work was<br />

conducted by personnel of <strong>the</strong> FS Region 3. By<br />

direction (Forest Service Manual 2676.2), survey<br />

data were recorded on standard <strong>for</strong>ms and maps<br />

following field observations. MTs were assigned<br />

using <strong>the</strong> survey results. Occupancy and reproductive<br />

status of <strong>Mexican</strong> spotted owls were<br />

summarized by MT at <strong>the</strong> close of each fiscal<br />

year (30 September). These summaries did not<br />

include all in<strong>for</strong>mation recorded on field <strong>for</strong>ms,<br />

such as that used to determine occupancy,<br />

reproductive status, or spatial coordinates of owl<br />

locations. Original data recorded on field <strong>for</strong>ms<br />

were necessary <strong>for</strong> verifying occupancy and<br />

reproductive status. Without this type of verification,<br />

reliability of previous assignments cannot<br />

be demonstrated. Fur<strong>the</strong>r, spatial coordinates<br />

compatible with a Geographic In<strong>for</strong>mation<br />

Volume II/Chapter 2<br />

2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

System (GIS) were required to document owl<br />

distribution and conduct habitat analyses at<br />

various spatial scales. Un<strong>for</strong>tunately, <strong>the</strong> original<br />

data and GIS-compatible coordinates were not<br />

entered into a computerized database. Thus, <strong>the</strong><br />

Team attempted to compile and create its own<br />

database (referred to as <strong>the</strong> Team database) from<br />

original data <strong>for</strong>ms and plotted locations to<br />

supplement summaries provided by <strong>the</strong> FS<br />

Region 3 Office (referred to as <strong>the</strong> FS database).<br />

Attempts to create <strong>the</strong> Team database met<br />

with limited success. Entering a portion of <strong>the</strong><br />

data <strong>for</strong>ms by a professional data-entry service<br />

failed because standard codes and recording<br />

instructions were not followed by data collectors.<br />

Fur<strong>the</strong>r, various details were not recorded and<br />

maps were missing. In a second attempt, <strong>the</strong><br />

Team <strong>for</strong>med and trained a set of crews to visit<br />

FS Ranger Stations and Supervisor Offices to<br />

enter data from <strong>for</strong>ms and maps and to record<br />

Universal Transverse Mercator (UTM) coordinates<br />

of owl locations associated with surveys.<br />

The same in<strong>for</strong>mation was collected from land<br />

management agencies throughout <strong>the</strong> owl’s<br />

range <strong>for</strong> <strong>the</strong> four-year period (1990-1993).<br />

Year-end summaries from <strong>the</strong> FS Region 3 were<br />

compared to output from <strong>the</strong> Team database to<br />

verify similarity <strong>for</strong> <strong>the</strong> period 1990-1993.<br />

We found many discrepancies between <strong>the</strong><br />

Team and FS Region 3 databases, <strong>the</strong> greatest<br />

being fewer territories with complete reproductive<br />

in<strong>for</strong>mation in <strong>the</strong> Team’s database. The FS<br />

database contained 1,984 records, whereas <strong>the</strong><br />

Team’s database contained only 377 observations<br />

where valid estimates of reproductive output<br />

(number of fledged young) were available. A<br />

valid estimate of reproductive output was one <strong>for</strong><br />

which appropriate protocols (see Forsman 1983,<br />

USDA Forest Service 1990) substantiated <strong>the</strong><br />

estimate or <strong>for</strong> which young were reported.<br />

Thus, we were able to locate only a fraction of<br />

<strong>the</strong> data presumably available.<br />

When <strong>the</strong> two databases were merged, we<br />

found 11 records in <strong>the</strong> Team database that had<br />

no corresponding record in <strong>the</strong> FS database.<br />

Fur<strong>the</strong>r, <strong>the</strong> FS database has no fecundity<br />

estimate <strong>for</strong> 45 records <strong>for</strong> which <strong>the</strong> Team’s<br />

database had a valid fecundity estimate. Of <strong>the</strong><br />

321 records that matched, 271 records had<br />

identical values <strong>for</strong> reproductive output. Eight of


<strong>the</strong> remaining records had greater values in <strong>the</strong><br />

Team database, whereas 42 had greater values in<br />

<strong>the</strong> FS database. For <strong>the</strong> 321 matching records,<br />

<strong>the</strong> FS database had a significantly greater<br />

fledgling rate (P < 0.001) than <strong>the</strong> Team database.<br />

The expected bias was that <strong>the</strong> FS database<br />

would have a lower estimate of reproductive<br />

output because many of <strong>the</strong> records with zero<br />

young reported were not properly substantiated<br />

by mousing data as was done in <strong>the</strong> Team<br />

database. However, changes by <strong>the</strong> FS occurred<br />

during summarization procedures, among <strong>the</strong>m<br />

selected territories with zero young recorded<br />

were changed to unconfirmed (Keith Fletcher,<br />

FS, Albuquerque, NM, pers. comm.). There<strong>for</strong>e,<br />

<strong>the</strong> FS database has a higher estimate of reproductive<br />

output because a number of territories<br />

with valid values of zero were eliminated from<br />

<strong>the</strong> calculation. This would introduce an overestimate<br />

of reproductive output. Whe<strong>the</strong>r <strong>the</strong> two<br />

biases in <strong>the</strong> FS database cancel each o<strong>the</strong>r is<br />

unknown.<br />

A potential problem with <strong>the</strong> Team database<br />

was <strong>the</strong> incomplete availability of data<br />

<strong>for</strong>ms; many of <strong>the</strong> needed <strong>for</strong>ms were missing.<br />

The last <strong>for</strong>m used to complete <strong>the</strong> estimate of<br />

reproduction <strong>for</strong> many of <strong>the</strong> MTs may not have<br />

been available <strong>for</strong> data entry. As a result, <strong>the</strong><br />

Team database has a lower estimate of reproduction<br />

than <strong>the</strong> FS database. The Team’s ef<strong>for</strong>t to<br />

validate <strong>the</strong> FS estimates of fecundity does not<br />

accomplish this objective. Ra<strong>the</strong>r, <strong>the</strong> Team<br />

database appears too incomplete to be useful <strong>for</strong><br />

estimating fecundity. There<strong>for</strong>e, we have used<br />

estimates of fecundity from <strong>the</strong> FS database with<br />

<strong>the</strong> understanding that <strong>the</strong>se estimates may be<br />

biased <strong>for</strong> two reasons: inadequate validation of<br />

<strong>the</strong> fecundity and occupancy results, and nonrandom<br />

sampling. The benefit of <strong>the</strong> Team<br />

obtaining raw data from <strong>the</strong> FS is that this ef<strong>for</strong>t<br />

has demonstrated severe problems with <strong>the</strong><br />

handling of data collection, data management,<br />

and data analysis <strong>for</strong> <strong>the</strong> monitoring program.<br />

By discovering <strong>the</strong>se problems, <strong>the</strong> problems are<br />

correctable in <strong>the</strong> future.<br />

Volume II/Chapter 2<br />

3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

LIFE LIFE HISTORY HISTORY PARAMETERS<br />

PARAMETERS<br />

PARAMETERS<br />

An organism’s life history is <strong>the</strong> combination<br />

of birth and death processes exhibited in its<br />

natural environment (Partridge and Sibley<br />

1991). The optimal trade-off in survival and<br />

reproduction, resulting in fitness, should be<br />

maximized by <strong>the</strong> life history favored by natural<br />

selection. Under optimality <strong>the</strong>ory, <strong>the</strong><br />

organism’s life history is a finely tuned result of<br />

adaptations to its environment; major perturbations<br />

to an organism’s environment could<br />

eventually lead to its extinction. However,<br />

behavioral plasticity may allow organisms to<br />

adjust life-history strategies to current environmental<br />

conditions (Hansen and Urban 1992).<br />

When examining recovery of a species after past<br />

and present environmental perturbations of<br />

varying magnitudes, one must question whe<strong>the</strong>r<br />

that species will perish or persist. There<strong>for</strong>e,<br />

examining <strong>the</strong> life history characteristics of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl is paramount to understanding<br />

how it responds to changes in its<br />

environment.<br />

An organism’s life history can be quantitatively<br />

described in terms of age- and sex-specific<br />

survival, age- and sex-specific fecundity rates,<br />

longevity, and age at first reproduction (Stearns<br />

1992). In <strong>the</strong> following section, we outline<br />

current estimates of life history parameters <strong>for</strong><br />

<strong>the</strong> <strong>Mexican</strong> spotted owl, using a variety of<br />

estimators calculated with different sources of<br />

data. These parameters can be used in two ways<br />

to make inferences about <strong>Mexican</strong> spotted owl<br />

populations: (1) to estimate <strong>the</strong> finite rate of<br />

population change (l), and (2) to examine<br />

components of fitness within populations (Roff<br />

1992).<br />

Age- Age- Age- and and Sex-Specific Sex-Specific Sex-Specific Survival<br />

Survival<br />

We estimated age- and sex-specific survival<br />

<strong>for</strong> <strong>Mexican</strong> spotted owls using mark-recapture<br />

estimators with data from banded owls in<br />

population studies, binomial survival estimators<br />

with data from radio-tagged owls studied at<br />

various locations, and survival estimators from<br />

data collected during <strong>the</strong> FS monitoring program.<br />

Where feasible, we recognized four age


classes: juveniles (J), first-year subadults (S1),<br />

second-year subadults (S2) and adults (A), all as<br />

described by Forsman (1981) and Moen et al.<br />

(1991). Juveniles (age, x = 0 years) were fledged<br />

young-of-<strong>the</strong>-year. First-year subadults (x = 1<br />

year), second-year subadults (x = 2 years), and<br />

adults (x ³ 3 years) have similar body plumage<br />

but were distinguished by different retrix characteristics<br />

(Moen et al. 1991). Not all studies<br />

differentiated <strong>the</strong> two subadult age classes, so<br />

<strong>the</strong>se two age classes were usually lumped into a<br />

single subadult (S) age class <strong>for</strong> <strong>the</strong> purpose of<br />

estimating survival. We distinguished between<br />

subadults and adults because subadults usually<br />

have lower reproductive rates than adults.<br />

Mark-recapture Mark-recapture Survival Survival Estimators Estimators from<br />

from<br />

Population Population Population Studies Studies<br />

Studies<br />

Mark-recapture estimators yield maximum<br />

likelihood estimates of apparent survival (f) and<br />

recapture (or resighting) probability (p), which<br />

are asymptotically unbiased, normally<br />

distributed, and have minimum variance<br />

(Lebreton et al. 1992). An important<br />

consideration with apparent survival is that<br />

1- f = death + permanent emigration. For<br />

apparent survival to accurately reflect true<br />

survival (S), permanent emigration over <strong>the</strong><br />

course of <strong>the</strong> study must be close to zero. Emigration<br />

is difficult to quantify, requiring <strong>the</strong> use<br />

of large samples of radio-marked birds.<br />

Ano<strong>the</strong>r limitation of <strong>the</strong> mark-recapture<br />

data is that only territorial birds are marked,<br />

because nonterritorial birds (“floaters”) do not<br />

respond to <strong>the</strong> capture and sighting methods<br />

employed. Even though marked juveniles enter<br />

<strong>the</strong> floater population, <strong>the</strong>y are not recaptured<br />

until <strong>the</strong>y become territorial, which may not<br />

happen because <strong>the</strong>y emigrate from <strong>the</strong> study<br />

area. Floaters that never become territorial are<br />

never included in <strong>the</strong> data to estimate survival,<br />

even if <strong>the</strong>y remain on <strong>the</strong> study area. As a result,<br />

juvenile and subadult survival estimates produced<br />

from birds banded as juveniles are biased<br />

low because of <strong>the</strong> entry into <strong>the</strong> floater population<br />

and/or emigration from <strong>the</strong> study area.<br />

However, estimates of survival generated from<br />

subadults and adults marked as territorial birds,<br />

hence with negligible emigration, are unbiased if<br />

Volume II/Chapter 2<br />

4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

inference is only to territorial birds. Thus,<br />

inferences from <strong>the</strong> mark-recapture data only<br />

apply to <strong>the</strong> territorial population because only<br />

birds from <strong>the</strong> territorial population are marked.<br />

We examined mark-recapture data from<br />

color-banded <strong>Mexican</strong> spotted owls derived from<br />

<strong>the</strong> three population study areas (Figure 2.1):<br />

(1) a 484-km 2 (187-mi 2 ) area located on <strong>the</strong><br />

Coconino National Forest (Upper Gila Mountains<br />

RU) denoted as <strong>the</strong> CSA (Gutiérrez et al.<br />

1994); (2) a 323-km 2 (125-mi 2 ) area located on<br />

<strong>the</strong> Gila National Forest (Upper Gila Mountains<br />

RU) denoted as <strong>the</strong> GSA (Gutiérrez et al. 1993);<br />

and (3) an area encompassing portions of <strong>the</strong><br />

Sky Island Mountains in sou<strong>the</strong>astern Arizona<br />

(Basin and Range - West RU) denoted as <strong>the</strong><br />

SISA (Duncan et al. 1993). A total of 148 adult,<br />

44 subadult, and 238 juvenile capture histories<br />

compiled during 3- and 4-year periods were<br />

utilized in <strong>the</strong> subsequent analyses (Table 2.1).<br />

Methods used to estimate survival from <strong>the</strong><br />

mark-recapture data are detailed in Burnham et<br />

al. (1994) and Lebreton et al. (1992). Two sets of<br />

parameters are estimated with <strong>the</strong>se models: f is<br />

<strong>the</strong> probability of a bird remaining alive and on<br />

<strong>the</strong> study area, and p is <strong>the</strong> probability that <strong>the</strong><br />

bird will be resighted after initial capture. Both f<br />

and p are indexed to provide specific estimates<br />

with respect to time, age, sex, and area. We<br />

employed three analyses: (1) goodness-of-fit<br />

testing to <strong>the</strong> Cormack-Jolly-Seber (CJS) models<br />

using computer program RELEASE (Burnham<br />

et al. 1987) and JOLLY (Pollock et al. 1990); (2)<br />

examination of a wide variety of models progressing<br />

from a biologically realistic global model<br />

to <strong>the</strong> simplest model using program SURGE<br />

(Lebreton et al. 1992); and (3) selection of <strong>the</strong><br />

most parsimonious model with Akaike’s In<strong>for</strong>mation<br />

Criteria (AIC) and likelihood ratio tests<br />

(LRT) (Lebreton et al. 1992).<br />

Goodness-of-fit tests of data to <strong>the</strong> CJS<br />

model include Test 2, which tests primarily <strong>for</strong><br />

lack of independence, and Test 3, which primarily<br />

tests <strong>for</strong> heterogeneity in survival and recapture<br />

probabilities (Burnham et al. 1987). We<br />

could only use Test 2 because of <strong>the</strong> short<br />

duration of <strong>the</strong> studies. Test 2 of <strong>the</strong> GSA and<br />

SISA data did not indicate lack of fit of <strong>the</strong> data<br />

(GSA: c 2 = 0.008, 1 df, P = 0.93; SISA:<br />

c 2 = 0.702, 1 df, P = 0.40). Test 2 was not


Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 2.1. 2.1. Location of Coconino (CSA), Gila (GSA), and Sky Islands (SISA) <strong>Mexican</strong> spotted owl<br />

population study areas in Arizona and New Mexico.<br />

FPO<br />

Table Table 2.1. 2.1. Time periods and number of capture histories from three study areas used in estimating<br />

<strong>Mexican</strong> spotted owl survival.<br />

Age Age Class Class When When B BBanded<br />

B anded<br />

Adults dults<br />

Subadults ubadults<br />

Juv uv uveniles uv eniles<br />

Coconino Coconino S SStudy<br />

S tudy Ar Area Ar ea<br />

(CSA)<br />

(CSA)<br />

1991-1993<br />

Gila ila S SStudy<br />

S tudy Ar Area Ar ea<br />

(GSA)<br />

(GSA)<br />

1991-1993<br />

Sky ky II<br />

Island I sland S SStudy<br />

SS<br />

tudy Ar Area Ar ea<br />

(SISA)<br />

(SISA)<br />

1990-1993<br />

41 52 55<br />

18 15 11<br />

95 84 59<br />

5


computable <strong>for</strong> <strong>the</strong> CSA because only one bird<br />

released in 1991 was not recaptured in 1992 but<br />

was captured in 1993.<br />

For model selection, <strong>the</strong> GSA and CSA<br />

data sets were modeled toge<strong>the</strong>r because <strong>the</strong>y<br />

employed similar designs and methodologies<br />

(Gutiérrez et al. 1993, 1994), both occur in <strong>the</strong><br />

Upper Gila Mountains RU, and we wanted to<br />

improve <strong>the</strong> precision of survival estimates by<br />

combining <strong>the</strong> data into one analysis. The<br />

procedures used allowed testing <strong>the</strong> assumption<br />

that survival and recapture rates were <strong>the</strong> same<br />

<strong>for</strong> <strong>the</strong>se two study areas. The SISA data set was<br />

modeled separately from <strong>the</strong> CSA and GSA data<br />

sets because it was conducted by a separate set of<br />

investigators using somewhat different protocols,<br />

and because this study occurred in <strong>the</strong> Basin and<br />

Range - West RU. In modeling <strong>the</strong> GSA and<br />

CSA data, we considered <strong>the</strong> global model to be<br />

{f g*s*a , p g*s*a } (AIC = 249.05, K = 28 parameters);<br />

it included study area (or group, g), sex (s), and<br />

age (a) effects. We did not include time effects<br />

because only two recapture periods were present<br />

in <strong>the</strong> data. The selected model<br />

{f J*g ,f S,A(GSA),AM(CSA) , f AF(CSA) , p S1 , p S2,A } (AIC =<br />

223.24, K = 6) included (1) separate estimates of<br />

juvenile (J) survival <strong>for</strong> each study area, (2) a<br />

single estimate <strong>for</strong> subadults (S) and adults (A)<br />

of both sexes on <strong>the</strong> GSA combined with subadults<br />

of both sexes and male adults (AM) on <strong>the</strong><br />

CSA, and (3) a single estimate <strong>for</strong> adult females<br />

(AF) on <strong>the</strong> CSA. No significant study area<br />

effects in terms of recapture probabilities (p)<br />

were found but <strong>the</strong>re was an age effect (S1 ¹ S2<br />

and A age classes). However, <strong>the</strong> estimate of f<br />

<strong>for</strong> CSA females was 1.000 ( se (f) = 0.000),<br />

which was unsuitable <strong>for</strong> modeling purposes.<br />

Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

O<strong>the</strong>r than indicating that adult females may<br />

^<br />

have higher survival than males, <strong>the</strong> f <strong>for</strong> adult<br />

females on <strong>the</strong> CSA was probably an artifact of<br />

limited number of capture occasions. Such a<br />

result is possible from a study with only three<br />

capture occasions. There<strong>for</strong>e, we used <strong>the</strong> nearest<br />

model to <strong>the</strong> selected model; i.e., <strong>the</strong> model<br />

with <strong>the</strong> next lowest AIC. This model {f , f , J*g A,S<br />

p , p } (AIC = 229.53,<br />

S1 A,S2<br />

K = 5) included separate estimates of juvenile<br />

survival <strong>for</strong> <strong>the</strong> two study areas and a single<br />

estimate <strong>for</strong> subadults and adults, with no sex<br />

effects (Table 2.2). Recapture probabilities were<br />

modeled <strong>the</strong> same as <strong>the</strong> previous model.<br />

Estimated differences in juvenile survival<br />

between <strong>the</strong> two study areas may represent<br />

differences in survey ef<strong>for</strong>t around <strong>the</strong> study<br />

areas and differences in study area size ra<strong>the</strong>r<br />

than real differences in <strong>the</strong> rates. For more<br />

explanatory details, refer to <strong>the</strong> Population<br />

Trends section.<br />

We could not obtain reliable estimates from<br />

<strong>the</strong> SISA. For all <strong>the</strong> preliminary models examined,<br />

survival estimates became bounded at <strong>the</strong><br />

upper limit of 1 indicating problems with <strong>the</strong><br />

estimation procedure. The selected model (f, p)<br />

^<br />

had a single survival estimate (f = 1.0000,<br />

se ^ ^<br />

(f) = 0.0000) and a single estimate of recapture<br />

probability (p ^ = 0.2366, se ^ ^ (p) = 0.0344)<br />

with no age, sex, or time effects on ei<strong>the</strong>r estimate.<br />

Such a conclusion is not biologically<br />

realistic; and given that <strong>the</strong> protocol used in this<br />

study did not stress resighting of marked birds<br />

(as evidenced by <strong>the</strong> low estimate of p), we do<br />

not feel <strong>the</strong>se results are useful.<br />

Table Table 2.2. 2.2. Apparent survival (f), recapture probability (p), and <strong>the</strong>ir sampling standard errors<br />

estimated <strong>for</strong> <strong>Mexican</strong> spotted owls between 1991 and 1993 on <strong>the</strong> Gila Study Area (GSA), New<br />

Mexico, and <strong>the</strong> Coconino Study Area (CSA), Arizona.<br />

Age-class Age-class (S (Study (S tudy ar area) ar ea)<br />

Juv uv uv uvenile uvenile<br />

enile (GSA)<br />

(GSA)<br />

Juv uv uvenile uv enile (CSA)<br />

(CSA)<br />

^ ^<br />

^ ^ ^ ^<br />

^<br />

0.2861<br />

se se ( (f) (<br />

0.0366<br />

0.0785<br />

Subadult ubadult & & A AAdult<br />

A dult (GSA (GSA & & CSA)<br />

CSA) 0.8889 0.0269 0.9818<br />

aRecapture probabilities <strong>for</strong> juveniles represent probability of recapture as an S1 or S2 individual.<br />

f<br />

0.0643<br />

6<br />

p<br />

0.7147 a<br />

0.7147 a<br />

^ se se ( (p) (<br />

0.1524<br />

0.1524<br />

0.0176


Binomial Binomial Survival Survival Estimators Estimators from<br />

from<br />

Radio-tracking Radio-tracking Studies<br />

Studies<br />

Binomial estimators allow estimation of<br />

true survival (S) where 1 - S = mortality. Estimates<br />

of S can be derived from radio-telemetry<br />

data and are preferable to estimates of apparent<br />

survival when permanent emigration occurs,<br />

radios do not influence survival, and sufficient<br />

numbers of radios are available <strong>for</strong> precise<br />

estimates. We estimated annual survival <strong>for</strong><br />

subadult and adult <strong>Mexican</strong> spotted owls based<br />

on 73 radio-marked individuals from six studies<br />

at four geographic locations (Table 2.3). We also<br />

used 22 juveniles from two independent studies,<br />

one in Colorado (R. Reynolds, unpub. data) and<br />

one in Utah (Willey 1992a, b). We selected <strong>the</strong><br />

Kaplan-Meier product limit estimator (Kaplan<br />

and Meier 1958) as modified by Pollock et al.<br />

(1989) to analyze <strong>the</strong> radio-telemetry data. This<br />

estimator allowed <strong>for</strong> staggered entry of individuals<br />

during <strong>the</strong> sampling period and <strong>the</strong> use<br />

of right-censored data (exact fates of individuals<br />

unknown due to radio failure) without incurring<br />

<strong>the</strong> biases due to censoring discussed by White<br />

and Garrott (1990). In addition, this estimator is<br />

nonparametric and, <strong>the</strong>re<strong>for</strong>e, does not require<br />

an underlying hazard function that must be<br />

ma<strong>the</strong>matically tractable (Pollock et al. 1989).<br />

Five important assumptions underlie <strong>the</strong><br />

Kaplan-Meier estimator: (1) individuals have<br />

been randomly sampled; (2) survival of <strong>the</strong><br />

marked animal is independent of o<strong>the</strong>r individuals;<br />

(3) attached radios do not influence survival;<br />

(4) censoring is random and unrelated to an<br />

individual’s fate; and (5) newly marked individuals<br />

have <strong>the</strong> same survival rate as previously<br />

marked individuals. We were unable to test <strong>the</strong>se<br />

assumptions because of low sample sizes and<br />

design of <strong>the</strong> studies. However, we felt that<br />

assumption (1) was probably not met because of<br />

<strong>the</strong> nature of <strong>the</strong> studies whereas assumptions<br />

(2)-(5) probably were met. However, controversy<br />

exists whe<strong>the</strong>r radios and <strong>the</strong>ir attachment<br />

(assumption 3) affect survival (Paton et al. 1991,<br />

Foster et al. 1992), although <strong>the</strong> studies reported<br />

here used tail-mounted radios ra<strong>the</strong>r than <strong>the</strong><br />

backpacks discussed by <strong>the</strong> references. This<br />

assumption cannot be tested with just radiotracking<br />

data.<br />

Volume II/Chapter 2<br />

7<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Adult and subadult age classes were pooled<br />

<strong>for</strong> analyses because of <strong>the</strong> small sample of<br />

subadults in <strong>the</strong> radio-marked samples and <strong>for</strong><br />

comparability with <strong>the</strong> mark-recapture estimates.<br />

Sex-specific estimates of adult and subadult<br />

survival were computed by pooling data across<br />

study areas. Pooling across study areas was<br />

necessary because of low sample sizes within<br />

each study area. In addition, data on both sexes<br />

from <strong>the</strong> Coconino National Forest were combined<br />

<strong>for</strong> comparison with <strong>the</strong> mark-recapture<br />

estimates from <strong>the</strong> CSA population study on<br />

that <strong>for</strong>est. The pooling of data across study<br />

areas and years prevented us from examining<br />

spatial and temporal variation in <strong>the</strong> Kaplan-<br />

Meier estimates of survival. In all Kaplan-Meier<br />

models, <strong>the</strong> annual sampling period started on 1<br />

June because most birds were radio-marked just<br />

after this date, this coincided with approximately<br />

<strong>the</strong> middle of <strong>the</strong> sampling period <strong>for</strong> <strong>the</strong><br />

population studies, and this was <strong>the</strong> “birth” date<br />

used <strong>for</strong> fledglings in population modeling<br />

(Franklin 1992). In all models, individuals<br />

tracked beyond <strong>the</strong> annual sampling period were<br />

recycled back to <strong>the</strong> beginning. For example, an<br />

individual that was radio-marked in July and<br />

<strong>the</strong>n had radio failure in August of <strong>the</strong> following<br />

year would be initially added in July, added again<br />

in <strong>the</strong> following June, and <strong>the</strong>n censored in<br />

August. In this way, <strong>the</strong> number of entries in <strong>the</strong><br />

model exceeded <strong>the</strong> actual number of individuals<br />

tagged and <strong>the</strong> only fates <strong>for</strong> an individual were<br />

to die or be censored from <strong>the</strong> analysis.<br />

Estimates of S <strong>for</strong> adults and subadults<br />

combined were close to identical <strong>for</strong> both sexes<br />

when study areas were pooled (Table 2.4). On<br />

<strong>the</strong> Coconino National Forest, Kaplan-Meier (S)<br />

and mark-recapture (f) estimates were compared<br />

<strong>for</strong> adult/subadult (both sexes combined) and<br />

juvenile age classes using a Wald test (Carroll<br />

and Ruppert 1988, Hosmer and Lemeshow<br />

1989):<br />

c 2<br />

(1) =<br />

^<br />

(S - f) 2<br />

var ^ ^ ^<br />

(S) + var ^ (f)<br />

Estimates from mark-recapture data were<br />

not significantly different from those estimated<br />

from <strong>the</strong> radio-telemetry data (c 2 = 0.950, 1 df,<br />

P = 0.330). Estimates <strong>for</strong> juvenile survival from<br />

^


8<br />

Table Table 2.3. 2.3. Description of radio-telemetry studies conducted on adult and subadult <strong>Mexican</strong> spotted owls in <strong>the</strong> Upper Gila (Nor<strong>the</strong>r AZ, Coconino<br />

NF, AZ), Basin and Range - East (Lincoln NF, NM), Sou<strong>the</strong>rn Rocky Mountains Colorado (Rocky Mts, CO), and Colorado Plateau (Zion NP, UT)<br />

RUs.<br />

FPO


<strong>the</strong> radio-telemetry data (Table 2.4) were not<br />

significantly different from estimates from markrecapture<br />

data collected on <strong>the</strong> CSA (c 2 = 0.752,<br />

1 df, P = 0.386) or <strong>the</strong> GSA (c 2 = 2.749, 1 df,<br />

P = 0.097).<br />

Survival Survival Estimators Estimators Estimators from from FS FS FS Region Region 3<br />

3<br />

Monitoring Monitoring Studies<br />

Studies<br />

The FS has inventoried <strong>Mexican</strong> spotted<br />

owl MTs since 1984, and conducted in<strong>for</strong>mal<br />

and <strong>for</strong>mal monitoring since 1989 (Table 2.5).<br />

Summaries of monitoring data were supplied by<br />

<strong>the</strong> FS (see Sources of In<strong>for</strong>mation section). For<br />

each MT, presence of owls (single male, single<br />

female, single unknown, unknown age and sex,<br />

or pair, or else unchecked) and result of reproduction<br />

(0, 1, 2, 3 young, unchecked, or unconfirmed)<br />

were noted.<br />

Persistence of a pair can be estimated from<br />

occupancy data ga<strong>the</strong>red during <strong>the</strong> FS spotted<br />

owl monitoring program. Pair persistence rate is<br />

defined here as <strong>the</strong> probability that a territory<br />

containing a pair of owls in one year will contain<br />

a pair in <strong>the</strong> succeeding year (but see discussion<br />

of biases below). An individual territory must be<br />

monitored both years to compute this statistic.<br />

The overall persistence rate was 80.5% <strong>for</strong> 824<br />

territories monitored with <strong>for</strong>mal and in<strong>for</strong>mal<br />

protocols (Table 2.6). Note that some territories<br />

had persistence estimates <strong>for</strong> <strong>the</strong>ir first year<br />

because some of <strong>the</strong>se MTs were surveyed as part<br />

of inventory <strong>the</strong> preceding year. No differences<br />

in <strong>the</strong> persistence rate across years were detected<br />

<strong>for</strong> <strong>the</strong> in<strong>for</strong>mal monitoring data (1989-1993;<br />

c 2 = 2.543, 4 df, P = 0.637), or <strong>the</strong> <strong>for</strong>mal<br />

monitoring data (1989-1993; c 2 = 3.233, 4 df,<br />

P = 0.520). We found no significant differences<br />

<strong>for</strong> year (c 2 = 4.2546, 4 df, P = 0.373), monitoring<br />

type (c 2 = 0.7133, 1 df, P = 0.398), or <strong>the</strong>ir<br />

interaction (c 2 = 1.3162, 4 df, P = 0.859) using<br />

a logistic regression model (Hosmer and<br />

Lemeshow 1989).<br />

The overall persistence rate based on pairs<br />

versus no pairs from 1989-1993 <strong>for</strong>mal and<br />

in<strong>for</strong>mal monitoring types (0.805, SE = 0.0138)<br />

was used to estimate adult survival (S a ). If pairs<br />

are assumed to remain on <strong>the</strong> same territory, lack<br />

of a pair on a territory that was previously<br />

occupied infers <strong>the</strong> death of one or both members<br />

of <strong>the</strong> pair. This is a difficult assumption to<br />

accept because pairs have been observed to<br />

change nest locations, but it allows some inference<br />

about adult survival. The dichotomy of<br />

pairs versus no pairs was used because this<br />

assumption is <strong>the</strong> simplest one possible <strong>for</strong><br />

estimating survival from persistence. Single birds<br />

(i.e., persistence of a single) could be included in<br />

<strong>the</strong> analysis to estimate survival, but would<br />

require an assumption that single males persisted<br />

on <strong>the</strong> territory at <strong>the</strong> same rate as single females.<br />

Ano<strong>the</strong>r bias is failure to detect a pair<br />

when both birds are present. Using <strong>the</strong> dichotomy<br />

of pairs versus no pairs, and assuming<br />

Table Table 2.4. 2.4. Estimates of true survival (S) <strong>for</strong> <strong>Mexican</strong> spotted owls based on radio-telemetry<br />

data.<br />

Volume II/Chapter 2<br />

FPO<br />

9<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 2.5. 2.5. Number of management territories checked one or more times during <strong>the</strong> <strong>Mexican</strong><br />

spotted owl monitoring program of FS Region 3.<br />

FPO<br />

Table Table 2.6. 2.6. Persistence of a <strong>Mexican</strong> spotted owl pair on a territory <strong>for</strong> <strong>for</strong>mal and in<strong>for</strong>mal monitoring<br />

data, with “persistence” defined as <strong>the</strong> probability a pair of owls will exist on a territory given<br />

that a pair was on <strong>the</strong> same territory <strong>the</strong> previous year. Data are from <strong>the</strong> monitoring program of FS<br />

Region 3.<br />

FPO<br />

10


that both adults survive at <strong>the</strong> same rate and<br />

independent of one ano<strong>the</strong>r, <strong>the</strong>n (S ) a 2 ^<br />

= 0.805,<br />

^ ^ ^<br />

or S = 0.897 (se (S ) = 0.0077). This estimate is<br />

a a<br />

very close to <strong>the</strong> observed estimate of apparent<br />

survival from <strong>the</strong> population studies (0.889).<br />

Two biases of opposite direction are possible<br />

<strong>for</strong> an estimate of adult survival based on<br />

pair persistence. A pair of owls may leave a<br />

territory <strong>for</strong> reasons o<strong>the</strong>r than <strong>the</strong> death of one<br />

member of <strong>the</strong> pair, resulting in an estimate of<br />

survival biased low. In contrast, one member of a<br />

pair may die, but <strong>the</strong> remaining member may be<br />

able to obtain ano<strong>the</strong>r mate and stay on <strong>the</strong><br />

territory. Ano<strong>the</strong>r possibility is that both members<br />

of a pair die, but <strong>the</strong> territory is occupied by<br />

a new pair. In <strong>the</strong>se situations, <strong>the</strong> estimate of<br />

survival is biased high. For this reason, we note<br />

<strong>the</strong> similarity of persistence-based survival (S a ) to<br />

mark-recapture estimates (f) but we relied solely<br />

on <strong>the</strong> mark-recapture estimates to estimate<br />

population trends.<br />

The 1989-1993 <strong>for</strong>mal and in<strong>for</strong>mal<br />

persistence data were also analyzed by recovery<br />

unit (Table 2.7) using a logistic regression model<br />

that included year (1989-93), recovery unit and<br />

<strong>the</strong> interaction between <strong>the</strong> two factors. No<br />

differences were found <strong>for</strong> year (c 2 = 5.9688,<br />

4 df, P = 0.202), recovery unit (c 2 = 2.6129,<br />

Volume II/Chapter 2<br />

^<br />

^<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

4 df, P = 0.624), or <strong>the</strong> interaction between <strong>the</strong><br />

two (c 2 = 13.1995, 15 df, P = 0.587). There<strong>for</strong>e,<br />

a reduced model with only recovery unit was<br />

analyzed; and again, no differences were found<br />

<strong>for</strong> recovery unit (c 2 = 7.0271, 4 df, P = 0.134).<br />

Survival Survival Estimates Estimates from from Band Band Return Return Data<br />

Data<br />

Records from <strong>the</strong> FWS Banding Laboratory<br />

were checked to determine if adequate banding<br />

data were available <strong>for</strong> estimation of survival. All<br />

records obtained from <strong>the</strong> Laboratory were part<br />

of studies reported above, so did not provide<br />

additional in<strong>for</strong>mation.<br />

Age- Age- and and Sex-specific Sex-specific Sex-specific Fecundity<br />

Fecundity<br />

Fecundity (m x ) can be defined as <strong>the</strong> mean<br />

annual number of live births of a given sex by a<br />

parent of that same sex over an interval of age<br />

(Caughley 1977), e.g., <strong>the</strong> number of female<br />

young per adult female. For <strong>Mexican</strong> spotted<br />

owls, we defined live births as <strong>the</strong> number of<br />

young fledging from <strong>the</strong> nest because <strong>the</strong> number<br />

of live births at hatching was not measured.<br />

To estimate fecundity, we initially used <strong>the</strong><br />

number of total young (e.g. of both sexes)<br />

fledged per pair as <strong>the</strong> response variable from <strong>the</strong><br />

Table Table 2.7. 2.7. Persistence of a pair of <strong>Mexican</strong> spotted owls on a territory <strong>for</strong> <strong>the</strong> five recovery units<br />

contained in <strong>the</strong> FS Region 3 monitoring database <strong>for</strong> 1989-1993, with “persistence” defined as <strong>the</strong><br />

probability a pair of owls will exist on a territory given that a pair was on <strong>the</strong> same territory <strong>the</strong> previous<br />

year. Survival rate is <strong>the</strong> probability that both members of <strong>the</strong> pair survived <strong>the</strong> year, and is <strong>the</strong><br />

square root of persistence.<br />

FPO<br />

11


two appropriate population studies (GSA and<br />

CSA) and <strong>the</strong> FS monitoring database. The<br />

third population study (SISA) was included in<br />

<strong>the</strong> analysis of <strong>the</strong> FS monitoring database<br />

because <strong>the</strong> methods used and territories monitored<br />

were identical. We analyzed <strong>the</strong> two<br />

population studies separately from <strong>the</strong> monitoring<br />

database to obtain fecundity estimates<br />

directly related to <strong>the</strong> mark-recapture estimates<br />

of survival. Estimates <strong>for</strong> number of young<br />

fledged were converted into fecundities by<br />

dividing <strong>the</strong> means in half and adjusting <strong>the</strong><br />

standard errors accordingly. This procedure<br />

assumed a 1:1 sex ratio at fledging and that sites<br />

where reproductive data were ga<strong>the</strong>red were<br />

independent across years.<br />

Fecundity Fecundity Estimates Estimates Estimates from from Population Population Population Studies<br />

Studies<br />

We used <strong>the</strong> general linear models procedure<br />

in SAS (PROC GLM; SAS Institute Inc.<br />

1985) to examine differences in time, age, and<br />

study areas <strong>for</strong> number of young fledged by male<br />

and female <strong>Mexican</strong> spotted owls. Average<br />

number of young fledged <strong>for</strong> all years differed<br />

between <strong>the</strong> GSA and CSA (F = 3.73; 1, 151 df;<br />

P = 0.055) but average number of young per<br />

year <strong>for</strong> combined study areas did not differ<br />

among years of study (F = 1.58; 2, 151 df;<br />

P = 0.210). Based on a priori linear contrasts,<br />

number of young fledged by first-year subadults<br />

was different from second-year subadults and<br />

adults combined <strong>for</strong> both males (F = 3.45; 1,<br />

151 df; P = 0.065) and females (F = 16.24; 1,<br />

151 df; P < 0.001; Table 2.8).<br />

Fecundity Fecundity Estimates Estimates from from FS FS Region Region 3<br />

3<br />

Monitoring Monitoring Database<br />

Database<br />

An objective of <strong>the</strong> FS Region 3 monitoring<br />

program was to monitor reproduction (Table<br />

2.9). Differences in mean reproduction across<br />

monitoring methods and years were tested with<br />

analysis of variance (ANOVA). Reproduction is a<br />

categorical variable, because only values of 0, 1,<br />

2, and 3 young are observed. However, ANOVA<br />

techniques are still appropriate because of <strong>the</strong><br />

large sample sizes involved (even though sample<br />

sizes are unequal); and hence, <strong>the</strong> cell means<br />

being approximately normally distributed as a<br />

Volume II/Chapter 2<br />

12<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

result. Ano<strong>the</strong>r possible procedure might be a<br />

log-linear analysis. However, <strong>the</strong> null hypo<strong>the</strong>sis<br />

of a log-linear model would be that <strong>the</strong> distributions<br />

are <strong>the</strong> same, compared to <strong>the</strong> null hypo<strong>the</strong>sis<br />

of ANOVA that <strong>the</strong> means are <strong>the</strong> same.<br />

Thus, a log-linear analysis may reject <strong>the</strong> null<br />

hypo<strong>the</strong>sis even when <strong>the</strong> mean fecundity rates<br />

do not differ. We did not use a log-linear analysis<br />

here because we are primarily interested in <strong>the</strong><br />

mean rates.<br />

All data <strong>for</strong> 1989-1993 were used to test <strong>for</strong><br />

differences between <strong>for</strong>mal and in<strong>for</strong>mal monitoring<br />

types (Table 2.9). The year effect was<br />

significant (F = 14.09; 4, 685 df; P < 0.001), but<br />

monitoring method (F < 0.001; 1, 685 df;<br />

P = 0.956) and <strong>the</strong> interaction (F = 1.23; 4,<br />

685 df; P = 0.295) were not significant. The<br />

average number of young fledged/pair across<br />

years was 1.006 (n = 695, SE = 0.037), giving a<br />

fecundity estimate of 0.503 (SE = 0.018).<br />

Differences in mean reproduction among<br />

recovery units with years included in <strong>the</strong> model<br />

was tested <strong>for</strong> <strong>for</strong>mal and in<strong>for</strong>mal monitoring<br />

data <strong>for</strong> 1989-1993. The year effect was significant<br />

(F = 3.03; 4, 672 df; P = 0.017), as were<br />

recovery unit (F = 12.55; 4, 672 df; P < 0.001),<br />

and <strong>the</strong>ir interaction (F = 2.86; 14, 672 df;<br />

P < 0.001). Mean reproductive rates are given in<br />

Table 2.10 <strong>for</strong> each year and recovery unit, plus<br />

<strong>the</strong> recovery unit mean across <strong>the</strong> years 1989-<br />

1993. Reproduction estimates from <strong>the</strong> FS<br />

monitoring database were compared to estimates<br />

from <strong>the</strong> population studies (CSA and GSA) <strong>for</strong><br />

<strong>the</strong> same two National Forests, i.e., Coconino<br />

and Gila. The ANOVA model included year<br />

(1991-1993), National Forest (Coconino, Gila),<br />

and monitoring method (population study<br />

versus <strong>for</strong>mal and in<strong>for</strong>mal monitoring), plus all<br />

<strong>the</strong> interactions of <strong>the</strong>se three factors. None of<br />

<strong>the</strong> factors was significant (P ³ 0.300) except <strong>the</strong><br />

interaction between <strong>for</strong>est and monitoring<br />

method (F = 5.24; 1, 267 df; P = 0.023). Table<br />

2.11 presents <strong>the</strong> four means that produced this<br />

interaction.<br />

Although <strong>the</strong> reasons behind this interaction<br />

are unknown, we speculate that two explanations<br />

are possible <strong>for</strong> this difference. First,<br />

differences in following <strong>the</strong> monitoring protocols<br />

between personnel in <strong>the</strong> Coconino and<br />

Gila National Forests may have led to <strong>the</strong>


13<br />

Table Table 2.8. 2.8. Sample size (n) and estimates (mean and SE) of number of young fledged/pair and fecundity (female young fledged/female) from <strong>the</strong> GSAa and CSAa Study Areas, 1991-1993.<br />

FPO


observed differences between <strong>the</strong> population<br />

studies and o<strong>the</strong>r monitoring systems. Monitoring<br />

approaches in <strong>the</strong> population studies are <strong>the</strong><br />

same because each was conducted by <strong>the</strong> same<br />

research group using standardized methods and<br />

experienced personnel. A second possibility is<br />

that <strong>the</strong> population study areas are not representative<br />

of <strong>the</strong> surrounding habitat <strong>for</strong> <strong>the</strong>ir respective<br />

<strong>for</strong>ests; and thus, <strong>the</strong> observed difference is<br />

real and caused by habitat differences on each<br />

<strong>for</strong>est. Alternatively, monitoring sites may not be<br />

representative of <strong>the</strong> <strong>for</strong>est because management<br />

territories were selected because of proposed<br />

timber sales (with better quality owl habitat) or<br />

historical locations (with easy access).<br />

Effects Effects of of Forest Forest Type Type on on Life<br />

Life<br />

History History History Traits<br />

Traits<br />

We examined <strong>the</strong> effects of <strong>for</strong>est type on<br />

both reproduction and persistence, which can be<br />

viewed as an indirect measure of survival. We<br />

used two sources of in<strong>for</strong>mation to examine <strong>the</strong><br />

effects of <strong>for</strong>est type on reproduction: <strong>the</strong> FS<br />

Region 3 monitoring database and data from<br />

Skaggs and Raitt (1988). For persistence, we<br />

used data from <strong>the</strong> FS Southwestern Region<br />

monitoring database. These were <strong>the</strong> only<br />

sources of data available to <strong>the</strong> Team which<br />

coupled habitat in<strong>for</strong>mation with data used to<br />

estimate life history traits.<br />

Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 2.9. 2.9. Number of <strong>Mexican</strong> spotted owl pairs checked (n), mean, and standard error (SE) <strong>for</strong><br />

number of young fledged per pair <strong>for</strong> 1989-1993, based on <strong>for</strong>mal and in<strong>for</strong>mal monitoring methods.<br />

FPO<br />

14<br />

Effects Effects on on Reproduction<br />

Reproduction<br />

Reproduction<br />

The FS Region 3 monitoring database<br />

included <strong>for</strong> each MT <strong>the</strong> percent (recorded to<br />

<strong>the</strong> nearest 25%) of <strong>the</strong> core area consisting of<br />

<strong>the</strong> following <strong>for</strong>est types: mixed conifer, pineoak,<br />

ponderosa pine, pinyon-juniper, oak,<br />

Arizona cypress, sycamore, o<strong>the</strong>r riparian, and<br />

unsuitable <strong>for</strong> owls. We computed <strong>the</strong> Pearson<br />

correlations (r) between each of <strong>the</strong>se <strong>for</strong>est type<br />

variables and number of young produced on an<br />

MT <strong>for</strong> each year, given that a pair of owls was<br />

present. Significant correlations were found <strong>for</strong><br />

mixed-conifer (r = -0.131, P < 0.001), pine-oak<br />

r = 0.084, P = 0.011), o<strong>the</strong>r riparian (r = 0.062,<br />

P = 0.064), and unsuitable (r = 0.098,<br />

P = 0.003), with none of <strong>the</strong> remaining variables<br />

significant (P > 0.154). This analysis suggested<br />

that <strong>the</strong> more mixed-conifer present, <strong>the</strong> lower<br />

<strong>the</strong> reproductive rate (a negative correlation),<br />

and <strong>the</strong> more unsuitable <strong>for</strong>est type, <strong>the</strong> greater<br />

<strong>the</strong> reproductive rate (a positive correlation).<br />

When <strong>the</strong> <strong>for</strong>est type variables were used in a<br />

step-wise regression to predict number of young<br />

fledged, mixed-conifer and unsuitable were both<br />

selected (P £ 0.017), while none of <strong>the</strong> remaining<br />

variables was included (P > 0.150). The<br />

regression explained only a very small amount<br />

(2.3%) of <strong>the</strong> variation in <strong>the</strong> number of young<br />

fledged, with <strong>the</strong> signs of both variables in <strong>the</strong><br />

opposite direction of what we expected based on<br />

radio-telemetry studies (Ganey and Dick 1995).


Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 2.10. 2.10. Number of <strong>Mexican</strong> spotted owl pairs checked (n), mean, and standard error (SE) <strong>for</strong><br />

number of young produced per pair <strong>for</strong> 1989-1993 by recovery unit, based on <strong>for</strong>mal and in<strong>for</strong>mal<br />

monitoring methods.<br />

FPO<br />

Table Table 2.11. 2.11. Sample size (n), mean, and standard error (SE) <strong>for</strong> <strong>the</strong> number of <strong>Mexican</strong> spotted owl<br />

young produced per territory <strong>for</strong> 1991-1993, based on population studies (CSA and GSAa ) and <strong>for</strong>mal<br />

and in<strong>for</strong>mal monitoring methods in <strong>the</strong> Coconino and Gila National Forests.<br />

FPO<br />

15


When we examined <strong>the</strong> database to determine<br />

<strong>the</strong> distribution of territories with mixed-conifer,<br />

we found that 276 territories with 100% mixedconifer<br />

cover occurred in <strong>the</strong> Basin and Range -<br />

East RU. These 276 territories were 79.3% of<br />

<strong>the</strong> territories that had 100% mixed-conifer.<br />

However, only 6 territories with


decreasing persistence, and ultimately fitness. As<br />

discussed previously <strong>for</strong> <strong>the</strong> fecundity analysis,<br />

<strong>the</strong> inequitable distribution of mixed conifer<br />

habitat across recovery units makes this analysis<br />

inappropriate.<br />

Conclusions<br />

Conclusions<br />

Environmental conditions may greatly<br />

affect reproduction and/or survival of nestlings<br />

through fledging and to adulthood. However,<br />

adult survival rates appear to be relatively constant<br />

across years, as suggested by high pair<br />

persistence rates. Such life history characteristics<br />

are common <strong>for</strong> K-selected species, <strong>for</strong> which<br />

populations remain relatively stable even though<br />

recruitment rates might be highly variable. With<br />

no recruitment, <strong>the</strong> population only declines at<br />

<strong>the</strong> rate of 1 minus adult survival, or <strong>the</strong> adult<br />

mortality rate.<br />

Undoubtedly, long-lived organisms experience<br />

a range of environmental conditions<br />

through time. Conditions that result in simultaneously<br />

low survival and reproduction can lower<br />

average population persistence rates dramatically,<br />

when examined over a period that is short<br />

relative to <strong>the</strong> organism’s life span. The converse,<br />

simultaneously high survival and reproduction,<br />

that would raise <strong>the</strong> expected population persistence<br />

dramatically, is also possible. However, <strong>the</strong><br />

magnitude of <strong>the</strong> effect of such a sudden decrease<br />

or increase on average persistence will<br />

naturally decline as <strong>the</strong> observation of a given<br />

population is extended in time, while <strong>the</strong> probability<br />

of detecting such events will increase with<br />

observation time. In addition, dispersal among<br />

subpopulations can greatly influence <strong>the</strong> persistence<br />

of relatively isolated populations (Keitt et<br />

al. 1995). Successful dispersal may also be a rare<br />

and time-dependent event or a density-dependent<br />

event. We currently have little in<strong>for</strong>mation<br />

on <strong>the</strong> frequency of immigration and its associated<br />

influence on population persistence. Thus,<br />

reliable conclusions on <strong>the</strong> persistence of <strong>Mexican</strong><br />

spotted owl populations must await additional<br />

study. Without reliable projections on <strong>the</strong><br />

owl’s persistence, we can only summarize conclusions<br />

about <strong>the</strong> owl’s survival and reproduction<br />

based on short-term but current knowledge.<br />

Volume II/Chapter 2<br />

17<br />

Survival<br />

Survival<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Annual survival rates of adult <strong>Mexican</strong><br />

spotted owls is ~0.8-0.9 based on short-term<br />

population and radio-tracking studies and<br />

longer-term monitoring studies. These annual<br />

survival estimates can be viewed as <strong>the</strong> probability<br />

of an individual surviving from one year to<br />

<strong>the</strong> next or as <strong>the</strong> proportion of individuals that<br />

will survive from one year to <strong>the</strong> next. A variety<br />

of different estimators of adult survival using<br />

different types and sets of data gave similar<br />

results. Juvenile survival is considerably lower<br />

(~0.06-0.29) than adult survival. Juvenile survival<br />

also appears more variable spatially, although<br />

this conclusion reflects only two population<br />

study areas and two radio-telemetry studies<br />

spanning two years or less.<br />

We strongly suspect that estimates of<br />

juvenile survival from <strong>the</strong> population studies are<br />

biased low because of (1) a high likelihood of<br />

permanent dispersal (emigration) from <strong>the</strong> study<br />

area, especially <strong>the</strong> smaller GSA, and (2) a lag of<br />

several years be<strong>for</strong>e marked juveniles reappear as<br />

territory holders, at which point <strong>the</strong>y are first<br />

detected <strong>for</strong> recapture. Concerning <strong>the</strong> first<br />

point, juvenile nor<strong>the</strong>rn spotted owls have a<br />

high dispersal capability (reviewed in Thomas et<br />

al. 1990). If <strong>Mexican</strong> spotted owl juveniles have<br />

a similar dispersal capability, we expect that a<br />

substantial portion of marked juveniles will<br />

emigrate from <strong>the</strong> respective study areas. However,<br />

estimates from <strong>the</strong> radio-telemetry study<br />

roughly corroborated <strong>the</strong> low estimates from <strong>the</strong><br />

population studies. Biases in <strong>the</strong> radio-telemetry<br />

estimates of juvenile survival can result if radios<br />

significantly affect <strong>the</strong>ir survival. Whe<strong>the</strong>r radios<br />

or <strong>the</strong>ir attachment affect survival of nor<strong>the</strong>rn<br />

spotted owls is debatable (Paton et al. 1991,<br />

Foster et al. 1992). Concerning <strong>the</strong> second<br />

point, Franklin (1992) found a lag of 1-4 years<br />

between <strong>the</strong> time when juvenile nor<strong>the</strong>rn<br />

spotted owls were banded and subsequently<br />

recaptured. If this process is similar <strong>for</strong> <strong>Mexican</strong><br />

spotted owls, <strong>the</strong>n <strong>the</strong> current population studies<br />

may be of insufficient duration to adequately<br />

estimate juvenile survival.<br />

In summary, our survival estimates are<br />

based primarily on studies of insufficient duration<br />

or studies not explicitly designed to estimate


survival. In most cases, <strong>the</strong> data were too limited<br />

to support or test <strong>the</strong> assumptions of <strong>the</strong> estimators<br />

used. However, <strong>the</strong> age- and sex-specific<br />

estimates of survival calculated here are useful at<br />

this point as qualitative descriptors of <strong>the</strong> lifehistory<br />

characteristics of <strong>Mexican</strong> spotted owls.<br />

That is, <strong>Mexican</strong> spotted owls exhibit high adult<br />

and relatively low juvenile survival. In this<br />

respect, <strong>Mexican</strong> spotted owl survival probabilities<br />

appear similar to nor<strong>the</strong>rn (see review in<br />

Burnham et al. 1994) and Cali<strong>for</strong>nia spotted<br />

owls (Noon and McKelvey 1992).<br />

Reproduction<br />

Reproduction<br />

Reproductive output of <strong>Mexican</strong> spotted<br />

owls, defined as <strong>the</strong> number of young fledged<br />

per pair, varies both spatially and temporally.<br />

<strong>Mexican</strong> spotted owls may have a higher average<br />

reproductive rate (1.001 fledged young per pair)<br />

than <strong>the</strong> Cali<strong>for</strong>nia (~0.712; Noon and<br />

McKelvey 1992) and <strong>the</strong> nor<strong>the</strong>rn spotted owl<br />

(~0.715; Thomas et al. 1990). Both of <strong>the</strong> o<strong>the</strong>r<br />

subspecies exhibit temporal fluctuations in<br />

reproduction similar to <strong>the</strong> <strong>Mexican</strong> subspecies.<br />

Effects Effects of of Forest Forest Type<br />

Type<br />

We feel <strong>the</strong> data collected by Skaggs and<br />

Raitt (1988) were more appropriate <strong>for</strong> examining<br />

<strong>the</strong> effects of <strong>for</strong>est type on reproduction<br />

than our analysis of <strong>the</strong> monitoring data. Their<br />

study was designed to look at <strong>the</strong> relationship<br />

between <strong>for</strong>est type, density and reproduction<br />

whereas <strong>the</strong> monitoring program was not.<br />

However, <strong>the</strong> Skaggs and Raitt data lacked<br />

sufficient statistical power to detect differences<br />

in reproductive output between <strong>the</strong> three <strong>for</strong>est<br />

types. This problem arose primarily because<br />

significantly fewer spotted owls were found in<br />

<strong>the</strong> pinyon-juniper and pine <strong>for</strong>est types than in<br />

mixed-conifer <strong>for</strong>ests (Ward et al. 1995). The<br />

primary problem in using data on <strong>for</strong>est types<br />

from MTs is that <strong>the</strong> data are confounded <strong>for</strong><br />

making <strong>the</strong> desired comparisons. Fur<strong>the</strong>r, <strong>the</strong><br />

core delineations were subjective. In addition,<br />

<strong>for</strong>est types included within MT boundaries may<br />

not have been used by <strong>the</strong> owls <strong>for</strong> which <strong>the</strong><br />

given MT was established. For <strong>the</strong>se reasons, we<br />

feel that definitive data linking <strong>Mexican</strong> spotted<br />

Volume II/Chapter 2<br />

18<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

owl reproduction and <strong>for</strong>est type are still lacking,<br />

even at <strong>the</strong> coarse-grained scale that we examined.<br />

Fur<strong>the</strong>r, <strong>for</strong>est type affects more than just<br />

reproduction; so additional data are needed to<br />

evaluate how <strong>for</strong>est type affects survival, and<br />

ultimately fitness.<br />

POPULATION POPULATION POPULATION TRENDS<br />

TRENDS<br />

We estimated trends in <strong>Mexican</strong> spotted<br />

owl populations two ways: as <strong>the</strong> finite rate of<br />

population change, lambda (l), and as trends in<br />

<strong>the</strong> rate of occupancy of spotted owl territories.<br />

Finite Finite Finite Rate Rate of of Population Population Population Change<br />

Change<br />

Lambda was computed <strong>for</strong> female <strong>Mexican</strong><br />

spotted owls from <strong>the</strong> age-specific survival and<br />

fecundity rates obtained from two population<br />

study areas, <strong>the</strong> GSA and CSA (see Life History<br />

Parameters section). Lambda is a useful metric<br />

because it measures both <strong>the</strong> direction and<br />

magnitude of change in population trends.<br />

Direction of population trends can be characterized<br />

as stationary (l = 1), declining (l < 1) or<br />

increasing (l > 1). The magnitude in change is<br />

expressed as <strong>the</strong> annual rate of change, R, where<br />

R = l - 1 <strong>for</strong> a birth-pulse population. From a<br />

population management perspective, <strong>the</strong> statistical<br />

hypo<strong>the</strong>sis is l < 1 versus <strong>the</strong> null hypo<strong>the</strong>sis<br />

that <strong>the</strong> population is ei<strong>the</strong>r stationary or increasing<br />

(l ³ 1). This is a one-sided test of <strong>the</strong><br />

<strong>for</strong>m:<br />

^<br />

1 - l<br />

^<br />

Z = ,<br />

se (l)<br />

where Z is normally distributed with m = 0,<br />

s 2 = 1.<br />

We used a Leslie matrix (Leslie 1945) as<br />

modified by Usher (1972) to compute estimates<br />

of l based solely on <strong>the</strong> estimates of age-specific<br />

fecundity and survival probabilities obtained<br />

from <strong>the</strong> two study areas. The <strong>for</strong>m of <strong>the</strong> matrix<br />

followed Usher (1972):<br />

f 0 m 1 f 1 m 2<br />

f 0 f 1<br />

,


where f 0 = juvenile survival, f 1 = survival <strong>for</strong> <strong>the</strong><br />

subadult and adult age classes combined,<br />

m 1 = S1 fecundity, and m 2 = fecundity <strong>for</strong> S2<br />

and adult age classes combined. The <strong>for</strong>m of <strong>the</strong><br />

matrix assumed a birth-pulse population with a<br />

post-breeding census and a projection interval of<br />

one year (Noon and Sauer 1992). Lambda was<br />

estimated as <strong>the</strong> dominant eigenvalue associated<br />

with <strong>the</strong> right eigenvector derived from <strong>the</strong><br />

matrix using power analysis (Caswell 1989). The<br />

se(l) was estimated using <strong>the</strong> delta method<br />

(Seber 1982, Alvarez-Buylla and Slatkin 1994),<br />

which included <strong>the</strong> sampling covariances <strong>for</strong> <strong>the</strong><br />

estimated survival probabilities.<br />

Parameter estimates used to compute l<br />

(Table 2.12) were taken from only <strong>the</strong> two<br />

population studies because all of <strong>the</strong> required<br />

parameters were estimated from data within <strong>the</strong><br />

same spatial and temporal scales. Survival estimates<br />

were based on <strong>the</strong> mark-recapture estimators.<br />

Estimates of l were computed separately<br />

<strong>for</strong> <strong>the</strong> two studies because of significant differences<br />

in fecundity and juvenile survival between<br />

<strong>the</strong> two areas.<br />

We obtained estimates of l which were<br />

significantly lower than 1 <strong>for</strong> <strong>the</strong> GSA but were<br />

not significantly different from 1 <strong>for</strong> <strong>the</strong> CSA<br />

(Table 2.13). These estimates of l represent<br />

trends in populations <strong>for</strong> only <strong>the</strong> places and<br />

times of study, and are based on only 3 years of<br />

data collection. We believe <strong>the</strong> estimate of l<br />

from GSA is 0.288). Presumably,<br />

<strong>the</strong> best quality territories were incorporated<br />

into <strong>the</strong> <strong>for</strong>mal monitoring system with a<br />

pair occupancy rate of 67.6% versus 61.0% <strong>for</strong>


Table Table 2.12.<br />

2.12.<br />

areas.<br />

Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table 2.12. Parameters used to estimate l <strong>for</strong> <strong>Mexican</strong> spotted owls on <strong>the</strong> GSA a and CSA a study<br />

FPO<br />

Table Table 2.13. 2.13. Estimates and standard errors of l <strong>for</strong> female <strong>Mexican</strong> spotted owls on <strong>the</strong> CSAa and<br />

GSAa study areas. The Z statistic and probability level are <strong>for</strong> a one-sided test of <strong>the</strong> null hypo<strong>the</strong>sis of<br />

l < 1 versus <strong>the</strong> alternative hypo<strong>the</strong>sis l ³ 1.<br />

FPO<br />

20


in<strong>for</strong>mal monitoring; and hence, <strong>for</strong>mal monitored<br />

territories were more likely to be occupied.<br />

However, <strong>the</strong> <strong>for</strong>mal monitoring system looks<br />

more intensively at <strong>the</strong> sites <strong>for</strong> owls, so this<br />

difference in occupancy rate may be because of<br />

search procedures, which we suggest as <strong>the</strong> most<br />

likely scenario.<br />

We examined differences in occupancy rate<br />

(Table 2.15) with a logistic regression model that<br />

included recovery unit, monitoring type (in<strong>for</strong>mal<br />

and <strong>for</strong>mal), and <strong>the</strong> interaction of <strong>the</strong>se<br />

two variables <strong>for</strong> <strong>the</strong> years 1989-1993. Type of<br />

monitoring was not significant (c 2 = 0.0191,<br />

1 df, P = 0.890), but recovery unit<br />

(c 2 = 22.7979, 4 df, P < 0.001) and <strong>the</strong> interaction<br />

of recovery unit and type of monitoring<br />

(c 2 = 16.1657, 4 df, P = 0.003) were significant.<br />

As defined and used here, occupancy rate is<br />

a ra<strong>the</strong>r artificial parameter because <strong>the</strong> selection<br />

of territories <strong>for</strong> inclusion in <strong>the</strong> monitoring<br />

database depends on judgement of human<br />

observers. MT boundaries are set subjectively, so<br />

<strong>the</strong>y do not necessarily represent owl home<br />

ranges. Fur<strong>the</strong>r, MTs may encompass more than<br />

one pair (May et al. in press), although this<br />

possibility is not supposed to occur. Changes in<br />

occupancy rate probably correspond more with<br />

<strong>the</strong> addition of new MTs to <strong>the</strong> list of those<br />

already monitored by <strong>the</strong> FS, level of ef<strong>for</strong>t used<br />

Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 2.14. 2.14. Estimates of population parameters <strong>for</strong> <strong>Mexican</strong> spotted owls on <strong>the</strong> GSAa and CSAa study areas, and <strong>the</strong> value of <strong>the</strong> parameter that gives l = 1, provided that all o<strong>the</strong>r parameters remain<br />

at <strong>the</strong>ir original estimate.<br />

FPO<br />

21<br />

to monitor <strong>the</strong> MTs (i.e., number of MTs<br />

monitored that meet <strong>for</strong>mal monitoring protocols),<br />

and o<strong>the</strong>r administrative factors ra<strong>the</strong>r<br />

than true change in <strong>the</strong> owl population. As a<br />

complicating factor, MTs are not a random<br />

sample of all existing <strong>Mexican</strong> spotted owl<br />

territories. As can be inferred from Figure 2.2,<br />

<strong>the</strong> percent of MTs occupied has dropped since<br />

1989 because all new MTs added to <strong>the</strong> monitoring<br />

system are initially occupied. Habitat<br />

changes induced by <strong>for</strong>est alterations over <strong>the</strong><br />

next several decades will make some of <strong>the</strong><br />

currently occupied MTs unsuitable. Thus, we<br />

expect that occupancy rate will decline <strong>for</strong><br />

existing territories. New MTs will probably be<br />

added to compensate <strong>for</strong> loss of existing territories,<br />

so change in occupancy rate does not<br />

provide a valid inference about changes in <strong>the</strong><br />

owl population.<br />

Conclusions<br />

Conclusions<br />

We have little confidence in our estimates<br />

of population trends <strong>for</strong> <strong>the</strong> following reasons.<br />

First, accurate and precise estimates of l depend<br />

on <strong>the</strong> accuracy and precision of <strong>the</strong> parameter<br />

estimates used in <strong>the</strong> calculations. Estimates of<br />

juvenile survival may be biased low <strong>for</strong> reasons<br />

stated previously and <strong>the</strong> time over which<br />

parameters have been estimated is insufficient.


Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 2.15. 2.15. Number of <strong>Mexican</strong> spotted owl territories checked (n) and <strong>the</strong> percent of <strong>the</strong>m occupied<br />

by a pair of owls <strong>for</strong> 1989-1993, based on <strong>for</strong>mal and in<strong>for</strong>mal monitoring methods.<br />

FPO<br />

Second, <strong>the</strong> population studies from which<br />

parameter estimates were derived have not been<br />

conducted <strong>for</strong> a sufficiently long period to<br />

capture temporal variation. If one considers l as<br />

an average rate of change over <strong>the</strong> study period,<br />

<strong>the</strong>n three years is insufficient to adequately<br />

estimate l from both biological and statistical<br />

perspectives. Third, rates of change estimated<br />

through occupancy provide little in<strong>for</strong>mation on<br />

how <strong>Mexican</strong> spotted owl populations are<br />

changing <strong>for</strong> <strong>the</strong> reasons stated previously.<br />

However, <strong>the</strong> analysis does illustrate why rates of<br />

population change using occupancy are inadequate<br />

<strong>for</strong> estimating trends in <strong>Mexican</strong> spotted<br />

owl populations. If nothing else, we believe <strong>the</strong><br />

analytical procedures and framework that we<br />

have used throughout this section should provide<br />

a template <strong>for</strong> future research as well as<br />

indicate priorities <strong>for</strong> future research ef<strong>for</strong>ts.<br />

Theoretical Theoretical Problems Problems with with Current Current<br />

Current<br />

Monitoring Monitoring Procedures<br />

Procedures<br />

Much ef<strong>for</strong>t has been expended by personnel<br />

of FS Region 3 in collecting <strong>the</strong> monitoring<br />

data summarized here. Un<strong>for</strong>tunately, <strong>the</strong><br />

monitoring ef<strong>for</strong>t was inadequate <strong>for</strong> detecting<br />

important changes in <strong>the</strong> population dynamics<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl because <strong>the</strong> appropriate<br />

parameters were not measured.<br />

The primary goal of <strong>the</strong> monitoring program<br />

should be <strong>the</strong> detection of significant<br />

changes in population levels of <strong>the</strong> owl. Such<br />

22<br />

changes can be caused by change in survival rates<br />

or change in reproduction, or both. The current<br />

monitoring program only examines reproduction.<br />

Occupancy rate and pair persistence are<br />

logically flawed approximations to survival.<br />

Thus, with <strong>the</strong> current monitoring system,<br />

drastic changes in <strong>the</strong> owl population could<br />

occur and not be detected; or if detected, <strong>the</strong><br />

current monitoring system would not yield <strong>the</strong><br />

statistical rigor necessary to substantiate <strong>the</strong><br />

conclusion strongly enough to withstand <strong>the</strong><br />

criticism of opponents to <strong>the</strong> suggested finding.<br />

There<strong>for</strong>e, an improved monitoring system must<br />

be developed <strong>for</strong> future work (USDI 1995).


Volume II/Chapter 2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 2.2. 2.2. Changes in occupancy of <strong>for</strong>mal and in<strong>for</strong>mal monitoring territories in <strong>the</strong> FS Region 3<br />

monitoring database.<br />

FPO<br />

FPO<br />

23


Volume II/Chapter 2<br />

LITERATURE LITERATURE CITED<br />

CITED<br />

Alvarez-Buylla, E. R., and M. Slatkin. 1994.<br />

Finding confidence limits on population<br />

growth rates: three real examples revised.<br />

Ecology 75:255-260.<br />

Burnham, K. P., D. R. Anderson, and G. C.<br />

White. 1994. Estimation of vital rates of <strong>the</strong><br />

nor<strong>the</strong>rn spotted owl. Colo. Coop. Fish and<br />

Wildl. Res. Unit, Colorado State Univ., Fort<br />

Collins. 44pp.<br />

——., ——., ——., C. Brownie, and K. H.<br />

Pollock. 1987. Design and analysis methods<br />

<strong>for</strong> fish survival experiments based on releaserecapture.<br />

Am. Fisheries Soc. Monogr. 5:1-<br />

437.<br />

Carroll, R. J., and D. Ruppert. 1988. Trans<strong>for</strong>mation<br />

and weighting in regression. Chapman<br />

and Hall, New York, N.Y. 249pp.<br />

Caswell, H. 1989. Matrix population models.<br />

Sinauer Assoc., Sunderland, Mass. 328pp.<br />

Caughley, G. 1977. Analysis of vertebrate<br />

populations. John Wiley and Sons, New York,<br />

N.Y. 234pp.<br />

Duncan, R. B., S. M. Speich, and J. D. Taiz.<br />

1993. Banding and blood sampling study of<br />

<strong>Mexican</strong> spotted owls in sou<strong>the</strong>astern Arizona.<br />

Unpubl. report, USDA For. Serv., Coronado<br />

Nat. For., Tucson, Ariz. 21pp.<br />

Forsman, E. D. 1981. Molt of <strong>the</strong> spotted owl.<br />

Auk 98:735-742.<br />

——. 1983. Methods and materials <strong>for</strong> locating<br />

and studying spotted owls. USDA For. Serv.<br />

Gen. Tech. Rep. PNW-162. 8pp.<br />

Foster, C. C., E. D. Forsman, E. C. Meslow, G.<br />

S. Miller, J. A. Reid, F. F. Wagner, A. B. Carey,<br />

and J. B. Lint. 1992. Survival and reproduction<br />

of radio-marked adult spotted owls. J.<br />

Wildl. Manage. 56:91-95.<br />

Franklin, A. B. 1992. Population regulation in<br />

nor<strong>the</strong>rn spotted owls: <strong>the</strong>oretical implications<br />

<strong>for</strong> management. Pages 815-827 in D.<br />

R. McCullough and R. H. Barrett, eds.<br />

Wildlife 2001: populations. Elsevier Applied<br />

Science, New York, N.Y.<br />

Ganey, J. L. 1988. Distribution and habitat<br />

ecology of <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong>s in Arizona.<br />

M.S. Thesis, Nor<strong>the</strong>rn Arizona Univ., Flagstaff.<br />

229 pp.<br />

24<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

——., and J.L. Dick, Jr. 1995. Habitat relationships<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl. Pages xxx -<br />

xxxx in USDI. <strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Vol. 2. USDI Fish and Wildl.<br />

Serv., Albuquerque, N.M.<br />

Gutiérrez, R. J., D. R. Olson, and M. E.<br />

Seamans. 1993. Demography of two <strong>Mexican</strong><br />

spotted owl <strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong> populations.<br />

Humboldt State University. Arcata,<br />

Cali<strong>for</strong>nia. 16pp.<br />

——., ——., and ——. 1994. Demography of<br />

two <strong>Mexican</strong> spotted owl populations in<br />

Arizona and New Mexico. Rep. submitted to<br />

Humboldt State Univ. Foundation, Arcata,<br />

Calif. 29pp.<br />

Hansen, A. J., and D. L. Urban. 1992. Avian<br />

response to landscape pattern: <strong>the</strong> role of<br />

species’ life histories. Landscape Ecol. 7:163-<br />

180.<br />

Hosmer, D. W., and S. Lemeshow. 1989. Applied<br />

logistic regression. John Wiley and Sons,<br />

New York, N.Y. 307pp.<br />

Kaplan, E. L., and P. Meier. 1958. Nonparametric<br />

estimation from incomplete observations.<br />

J. Am. Statist. Assoc. 53:457-481.<br />

Keitt, T., A.B. Franklin, and D.L. Urban. 1995.<br />

Landscape analysis and metapopulation<br />

structure. Chapter 3 (16pp.) in USDI. <strong>Recovery</strong><br />

plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl. Vol. II.<br />

USDI Fish and Wildl. Serv., Albuquerque,<br />

N.M.<br />

Kroel, K. and P. J. Zwank. 1991. Home range<br />

and habitat use characteristics of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl in <strong>the</strong> sou<strong>the</strong>rn Sacramento<br />

Mountains, New Mexico. Unpubl. rep.,<br />

USDA For. Serv., Lincoln National Forest.<br />

New Mexico Coop. Fish and Wildl. Unit.<br />

Lebreton, J. D., K. P. Burnham, J. Clobert, and<br />

D. R. Anderson. 1992. Modeling survival and<br />

testing biological hypo<strong>the</strong>ses using marked<br />

animals: a unified approach with case studies.<br />

Ecol. Monogr. 62:67-118.<br />

Leslie, P. H. 1945. On <strong>the</strong> use of matrices in<br />

certain population ma<strong>the</strong>matics. Biometrika<br />

35:183-212.<br />

May, C.A., M.Z. Perry, R.J. Gutiérrez, M.E.<br />

Seamans, and D.R. Olson. In press. Feasibility<br />

of a random quadrat study design to estimate<br />

changes in density of <strong>Mexican</strong> spotted owls.<br />

U.S. For. Serv. Res. Pap.


Moen, C. A., A. B. Franklin, and R. J. Gutiérrez.<br />

1991. Age determination of subadult nor<strong>the</strong>rn<br />

spotted owls in northwest Cali<strong>for</strong>nia.<br />

Wildl. Soc. Bull. 19:489-493.<br />

Noon, B. R., and K. S. McKelvey. 1992. Stability<br />

properties of <strong>the</strong> spotted owl<br />

metapopulation in sou<strong>the</strong>rn Cali<strong>for</strong>nia. Pages<br />

187-206 in J. Verner, K. S. McKelvey, B. R.<br />

Noon, R. J. Gutiérrez, G. I. Gould, Jr., and T.<br />

W. Beck, eds. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status.<br />

USDA For. Serv. Gen. Tech. Rep. PSW-133.<br />

——., and J. R. Sauer. 1992. Population models<br />

<strong>for</strong> passerine birds: structure, parameterization,<br />

and analysis. Pages 441-464 in D. R.<br />

McCullough and R. H. Barrett, eds. Wildlife<br />

2001: Populations. Elsevier Applied Science,<br />

New York, N.Y.<br />

Partridge, L., and R. Sibly. 1991. Constraints in<br />

<strong>the</strong> evolution of life histories. Philos. Trans. R.<br />

Soc. Lond., Ser. B 332:3-13.<br />

Paton, P. W. C., C. J. Zabel, D. L. Neal, G. N.<br />

Steger, N. G. Tilghman, and B. R. Noon.<br />

1991. Effects of radio tags on spotted owls. J.<br />

Wildl. Manage. 55:617-622.<br />

Pollock, K. H., J. D. Nichols, C. Brownie, and J.<br />

E. Hines. 1990. Statistical inference <strong>for</strong><br />

capture-recapture experiments. Wildl.<br />

Monogr. 107:1-97.<br />

——., Winterstein, C. M. Bunck, and P. D.<br />

Curtis. 1989. Survival analysis in telemetry<br />

studies: <strong>the</strong> staggered entry design. J. Wildl.<br />

Manage. 53:7-15.<br />

Rinkevich, S.E., J.L. Ganey, W.H. Mow, F.P.<br />

Howe, F. Clemente, and J.F. Martinez-<br />

Montoya. 1995. <strong>Recovery</strong> units. Pages 36-51<br />

in USDI. <strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Vol. I. USDI Fish and Wildl.<br />

Serv., Albuquerque, N.M.<br />

Roff, D. A. 1992. The evolution of life histories:<br />

<strong>the</strong>ory and analysis. Chapman and Hall. New<br />

York, N.Y. 535pp.<br />

SAS Institute, Inc. 1985. SAS language guide <strong>for</strong><br />

personal computers, Version 6. SAS Institute,<br />

Cary, N.C. 429pp.<br />

Seber, G. A. F. 1982. Estimation of animal<br />

abundance. Second ed. C. Griffin, London,<br />

U.K. 654pp.<br />

Volume II/Chapter 2<br />

25<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Skaggs, R. W., and R. J. Raitt. 1988. A spotted<br />

owl inventory of <strong>the</strong> Lincoln National Forest,<br />

Sacramento Division, 1988. New Mexico<br />

Dep. Game and Fish, Santa Fe. 12pp.<br />

Stearns, S. C. 1992. The evolution of life histories.<br />

Ox<strong>for</strong>d Univ. Press, Ox<strong>for</strong>d, U.K. 249pp.<br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner. 1990. A<br />

conservation strategy <strong>for</strong> <strong>the</strong> spotted owl. U.S.<br />

Gov. Print. Off., Washington, D.C. 427pp.<br />

USDA Forest Service. 1990. Management<br />

guidelines and inventory and monitoring<br />

protocols <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl in <strong>the</strong><br />

Southwestern Region. Federal Register<br />

55:27278-27287.<br />

USDI. 1995. <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong><br />

<strong>Plan</strong>. Vol. 1. USDI Fish and Wildl. Serv.,<br />

Albuquerque, N.M.<br />

Usher, M. B. 1972. Developments in <strong>the</strong> Leslie<br />

matrix model. Pages 29-60 in J. N. R. Jeffers,<br />

ed. Ma<strong>the</strong>matical models in ecology.<br />

Blackwell Scientific Publ., Ox<strong>for</strong>d, U.K.<br />

Ward, J.P., Jr., S.E. Rinkevich, A.B. Franklin,<br />

and F. Clemente. 1995. Distribution and<br />

abundance. Chapter 1 (14pp.) in USDI.<br />

<strong>Recovery</strong> plan <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Vol. II. USDI Fish and Wildl. Serv.,<br />

Albuquerque, N.M.<br />

White, G. C., and R. A. Garrott. 1990. Analysis<br />

of wildlife radio-tracking data. Academic<br />

Press, San Diego, Calif. 383pp.<br />

Willey, D. W. 1992a. Semi-annual report: home<br />

range characteristics of <strong>Mexican</strong> spotted owls<br />

in <strong>the</strong> canyonlands geographic province;<br />

winter 1991-1992. Unpubl. rep., Utah Div.<br />

Wildl. Resour., Salt Lake City.<br />

——. 1992b. Movements and habitat ecology of<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong>s in sou<strong>the</strong>rn Utah.<br />

Final Rep. submitted to Utah Div. Wildl.<br />

Resour., Salt Lake City.<br />

——. 1993. Home-range characteristics and<br />

juvenile dispersal ecology of <strong>Mexican</strong> spotted<br />

<strong>Owl</strong>s in sou<strong>the</strong>rn Utah. Unpubl. rep., Utah<br />

Div. Wildl. Resour., Salt Lake City.


Unlike <strong>the</strong> nor<strong>the</strong>rn spotted owl (see<br />

Thomas et al. 1990), <strong>the</strong> range-wide population<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl is naturally fragmented<br />

into geographically distinct subpopulations.<br />

Thus far, we have discussed populations at<br />

<strong>the</strong> local scale within <strong>Recovery</strong> Units where<br />

individuals within subpopulations interact to<br />

some degree. Understanding population structure<br />

at larger scales is important, even though<br />

data are extremely limited at <strong>the</strong>se scales <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. Hanski and Gilpin (1991)<br />

identified two additional scales beyond <strong>the</strong> local<br />

scale: (1) a metapopulation scale where individuals<br />

infrequently move between subpopulations<br />

and typically cross unsuitable habitat to reach an<br />

adjacent subpopulation, and (2) a geographical<br />

scale where individuals have little likelihood of<br />

moving to most portions of <strong>the</strong> geographic range<br />

of <strong>the</strong>ir species. By definition, metapopulations<br />

are systems of local populations connected by<br />

dispersing individuals (Hanski and Gilpin<br />

1991). The metapopulation concept has been<br />

applied to <strong>the</strong> conservation of nor<strong>the</strong>rn<br />

(Schaeffer 1985) and Cali<strong>for</strong>nia (Noon and<br />

McKelvey 1992) spotted owl populations.<br />

However, on a geographical scale, <strong>the</strong> rangewide<br />

population of <strong>Mexican</strong> spotted owls may be<br />

composed of a several discrete metapopulations<br />

ra<strong>the</strong>r than of a single integrated<br />

metapopulation. Understanding to what degree<br />

<strong>Mexican</strong> spotted owl populations follow<br />

metapopulation dynamics is essential <strong>for</strong> managing<br />

<strong>the</strong> subspecies on all three scales.<br />

METAPOPULATION METAPOPULATION MODELS<br />

MODELS<br />

Theoretical work has proposed several<br />

models of metapopulations based on different<br />

mechanisms <strong>for</strong> persistence. These have been<br />

classified as <strong>the</strong> Levins model, source-sink, coresatellite,<br />

patchy, and non-equilibrium<br />

metapopulations (Harrison 1991). An important<br />

consideration when reviewing <strong>the</strong>se models is<br />

that empirical evidence <strong>for</strong> <strong>the</strong> existence of <strong>the</strong><br />

Volume II/Chapter 3<br />

CHAPTER CHAPTER 3: 3: Landscape Landscape Analysis<br />

Analysis<br />

and and and Metapopulation Metapopulation Structure<br />

Structure<br />

Tim Keitt, Alan Franklin, and Dean Urban<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

proposed mechanisms in natural populations is<br />

debatable (Doak and Mills 1994).<br />

The The Levins Levins Metapopulation Metapopulation Model<br />

Model<br />

Levins (1969) first introduced <strong>the</strong><br />

metapopulation concept with a simple model.<br />

This model incorporated three essential requirements<br />

<strong>for</strong> persistence of a subdivided population:<br />

(1) density-dependent dynamics of populations,<br />

(2) asynchronous dynamics of local<br />

subpopulations in that not all subpopulations<br />

have <strong>the</strong> same rate of change at <strong>the</strong> same time,<br />

and (3) dispersal between subpopulations.<br />

Dispersal is an essential component of this, and<br />

all o<strong>the</strong>r, metapopulation models and provides<br />

<strong>the</strong> mechanism <strong>for</strong> recolonizing areas where local<br />

subpopulations have died out.<br />

Source-sink Source-sink and and Core-satellite<br />

Core-satellite<br />

Metapopulation Metapopulation Metapopulation Models Models<br />

Models<br />

In source-sink models (Pulliam 1988),<br />

source areas with self-propagating (typically<br />

increasing) populations provide a flow of recruits<br />

to sink areas where populations are not selfreproducing<br />

(and may be declining). Without<br />

<strong>the</strong> net flow of immigrants from <strong>the</strong> source<br />

areas, populations in sinks would not persist.<br />

Source-sink mechanisms can be applied to<br />

contiguously distributed populations where<br />

habitat conditions can dictate whe<strong>the</strong>r a population<br />

segment is a source or a sink. This mechanism<br />

can also be applied to a metapopulation of<br />

discrete subpopulations where habitat or o<strong>the</strong>r<br />

conditions dictate whe<strong>the</strong>r such an area is a<br />

source or a sink.<br />

The core-satellite model builds on <strong>the</strong><br />

source-sink model by having a central core<br />

population which acts as a source surrounded by<br />

a number of smaller sink populations (Harrison<br />

1991). Persistence of <strong>the</strong> satellite populations<br />

depends upon <strong>the</strong> central source population.


Patchy Patchy Metapopulation Metapopulation Models<br />

Models<br />

Patchy populations in <strong>the</strong> context of this<br />

model are characterized by a high dispersal<br />

potential where dispersal takes place on a spatial<br />

scale greater than <strong>the</strong> local events causing subpopulation<br />

dynamics (Harrison 1991). The<br />

patchy metapopulation model is distinguished<br />

from <strong>the</strong> Levins model in that <strong>the</strong> average<br />

individual can belong to more than one subpopulation<br />

in its lifetime (excluding its natal<br />

subpopulation). The result, in effect, is a wellmixed<br />

version of <strong>the</strong> Levins model.<br />

Non-equilibrium Non-equilibrium Metapopulation<br />

Metapopulation<br />

Model<br />

Model<br />

Populations characterized by this model<br />

essentially follow <strong>the</strong> same dynamics as <strong>the</strong><br />

Levins model except that continuous<br />

recolonization of extinction-prone subpopulations<br />

has been disrupted. Essentially, this model<br />

represents metapopulations in decline even<br />

though one or more subpopulations may be<br />

stable. Instability of a <strong>for</strong>merly stable<br />

metapopulation may result from factors that<br />

affect one or more subpopulations or when<br />

contiguous populations are fragmented into<br />

discrete subpopulations. These factors can<br />

include natural (e.g., fire, disease) or anthropogenic<br />

(e.g., logging, urban development) disturbances.<br />

Volume II/Chapter 3<br />

DISPERSAL<br />

DISPERSAL<br />

In all metapopulation models, dispersal is a<br />

key component. Dispersal acts as a bridge<br />

between subpopulations at <strong>the</strong> metapopulation<br />

scale to provide immigrants to o<strong>the</strong>rwise isolated<br />

habitat patches. The fate of <strong>the</strong>se recruits depends<br />

on <strong>the</strong> status of <strong>the</strong> subpopulation. If <strong>the</strong><br />

habitat patch has been unoccupied, <strong>the</strong>n <strong>the</strong> new<br />

recruits “rescue” it from extinction. If <strong>the</strong> patch<br />

is not saturated with territorial breeders, <strong>the</strong><br />

recruits might bolster <strong>the</strong> breeding population.<br />

If <strong>the</strong> patch is saturated, <strong>the</strong> new recruits might<br />

persist as nonbreeding “floaters,” perhaps buffering<br />

<strong>the</strong> population against future fluctuations.<br />

2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Adult and subadult <strong>Mexican</strong> spotted owls<br />

appear to be relatively sedentary once <strong>the</strong>y<br />

come to occupy a given site. Juvenile <strong>Mexican</strong><br />

spotted owls, however, almost always disperse<br />

from <strong>the</strong>ir natal sites (Willey 1993, Hodgson<br />

and Stacey 1994, Reynolds and Johnson 1994).<br />

If metapopulation dynamics apply to <strong>Mexican</strong><br />

spotted owl populations, <strong>the</strong>se dynamics probably<br />

hinge on <strong>the</strong> flow of juvenile spotted owls<br />

between subpopulations.<br />

Dispersal of young <strong>Mexican</strong> spotted owls is<br />

poorly understood. Several studies have attempted<br />

to examine juvenile dispersal through<br />

radio-telemetry (Willey 1993, Hodgson and<br />

Stacey 1994, Reynolds and Johnson 1994); but<br />

sample sizes have been small. Total distances<br />

moved by 7 juveniles radio-marked in Utah<br />

(Willey 1993) ranged from 32 to 98 km (20-61<br />

miles) (median = 41.8 km [26 miles]). Four of<br />

<strong>the</strong>se juveniles moved back to within 8 km<br />

(5 miles) of <strong>the</strong>ir natal area just prior to <strong>the</strong><br />

breeding season. Hodgson and Stacey (1994)<br />

also reported 2 juveniles dispersing between <strong>the</strong><br />

San Mateo and Black Mountain Ranges of New<br />

Mexico, distances of 45 and 58 km (28-36<br />

miles). In addition, dispersing juveniles crossed<br />

large expanses of habitat typically considered<br />

unsuitable <strong>for</strong> resident spotted owls. Reynolds<br />

and Johnson (1994) radio-marked 6 juveniles,<br />

none of which were relocated beyond <strong>the</strong>ir natal<br />

site. The radio-telemetry data suggest that<br />

juvenile owls have <strong>the</strong> dispersal capability to act<br />

as recolonizers between many of <strong>the</strong> subpopulations<br />

and to provide genetic links between<br />

subpopulations. Thus far, none of <strong>the</strong> radiomarked<br />

juveniles have been recruited into a<br />

resident, territorial subpopulation.<br />

With source-sink and core-satellite<br />

metapopulation mechanisms, juvenile owls from<br />

source areas must not only reach isolated subpopulations<br />

but do so in sufficient numbers to<br />

stabilize sink populations. We examined straightline<br />

distances moved by 25 juveniles banded<br />

from 1991 through 1993 on <strong>the</strong> population<br />

study areas (CSA and GSA) which had been<br />

recaptured as territorial subadults and adults<br />

(R.J. Gutiérrez, D. Olsen, and M. Seamans,<br />

Humboldt State Univ., Arcata, CA, pers.<br />

comm.). From <strong>the</strong>se data, we generated a probability<br />

function by fitting an exponential curve


to distances moved and <strong>the</strong> cumulative probability<br />

that juveniles had moved at least those<br />

distance (Figure 3.1). This curve represents <strong>the</strong><br />

probability that a juvenile will disperse at least a<br />

given distance and be recruited into <strong>the</strong> territorial<br />

population. This curve can also be viewed in<br />

terms of proportions. For example, about 60%<br />

of <strong>the</strong> juveniles would be expected to disperse at<br />

least 10 km (6.2 miles) and be recruited into <strong>the</strong><br />

population. We provide this in<strong>for</strong>mation as an<br />

illustration of <strong>the</strong> type of data needed to assess<br />

<strong>the</strong> role of dispersal in metapopulation dynamics.<br />

The data we used were not designed to<br />

generate such a curve and may be artificially<br />

truncated because of lower search ef<strong>for</strong>t <strong>for</strong><br />

banded recruits in <strong>the</strong> larger areas surrounding<br />

each of <strong>the</strong> study areas. Such an effect can be<br />

responsible <strong>for</strong> <strong>the</strong> exponential nature of <strong>the</strong><br />

relationship in Figure 3.1 and underestimate true<br />

dispersal distances.<br />

APPLICATION APPLICATION OF OF LANDSCAPE<br />

LANDSCAPE<br />

LANDSCAPE<br />

ECOLOGY<br />

ECOLOGY<br />

Landscape ecology is concerned with <strong>the</strong><br />

development, implications, and dynamics of<br />

landscape pattern. We use <strong>the</strong> term “pattern”<br />

generously, but here we are concerned especially<br />

with three components of pattern: (1) <strong>the</strong> size<br />

distribution of habitat patches; (2) <strong>the</strong> spatial<br />

orientation of <strong>the</strong>se patches with respect to each<br />

o<strong>the</strong>r, that is, <strong>the</strong>ir juxtaposition; and (3) <strong>the</strong>ir<br />

mutual distance relationships as <strong>the</strong>se might<br />

influence spotted owl dispersal.<br />

With respect to metapopulation dynamics,<br />

landscape analysis is concerned with distance<br />

relationships among habitat patches as <strong>the</strong>se<br />

might affect <strong>the</strong> relative isolation of certain<br />

patches or regions of patches within <strong>the</strong> owl’s<br />

range where patches may correspond to distinct<br />

subpopulations. The issue of so-called patch<br />

“connectedness” is central to current ideas about<br />

metapopulation dynamics. By conventional<br />

definition, two patches are functionally connected<br />

if individuals can disperse between <strong>the</strong>m<br />

with a probability or frequency above some<br />

minimum threshold value. We will fur<strong>the</strong>r<br />

define connectedness to incorporate patch area<br />

as well as between-patch distance relationships<br />

Volume II/Chapter 3<br />

3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

(see below); thus, a patch may be highly connected<br />

if it is near several small patches or near<br />

one large patch. Specifically, we have two concerns<br />

about patch connectedness:<br />

1. Can we identify habitat patches that<br />

because of <strong>the</strong>ir area and position within<br />

<strong>the</strong> landscape might play a particularly<br />

important role in overall landscape<br />

connectedness? We expect large habitat<br />

patches to be important in this analysis.<br />

2. Are <strong>the</strong>re habitat patches that might be<br />

critical to landscape connectedness<br />

chiefly because of <strong>the</strong>ir spatial position,<br />

that is, despite <strong>the</strong>ir smaller size? These<br />

might function as dispersal conduits or<br />

small “stepping stones” within <strong>the</strong><br />

landscape.<br />

The goal of this set of questions is to identify<br />

habitat patches that might influence patterns in<br />

owl distribution well beyond <strong>the</strong>ir immediate<br />

location. In <strong>the</strong> case of a small stepping-stone<br />

patch, this importance might be despite its<br />

having ra<strong>the</strong>r modest owl populations. Thus,<br />

critical patches identified in <strong>the</strong> landscape<br />

analysis might warrant special consideration in<br />

<strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> even though <strong>the</strong>ir local<br />

populations might not elicit any special concern.<br />

Scale Scale and and Resolution Resolution in in Landscape Landscape<br />

Landscape<br />

Analysis<br />

Analysis<br />

Landscape analysis typically connotes ra<strong>the</strong>r<br />

large spatial scales. In our analyses, we have<br />

identified two spatial scales of interest. At a<br />

smaller scale we are concerned with areas such as<br />

a single National Forest, a scale defined in large<br />

part by administrative criteria. At a larger scale<br />

we are concerned with <strong>the</strong> geographic range of<br />

<strong>the</strong> owl as we defined it in this plan. We will<br />

refer to <strong>the</strong>se scales as “<strong>for</strong>est” and “rangewide.”<br />

Importantly, habitat data available at <strong>the</strong>se two<br />

scales is of very different resolution. Data at <strong>the</strong><br />

<strong>for</strong>est scale are of higher spatial resolution<br />

(smaller grain size) and often of higher in<strong>for</strong>mation<br />

content as well, while data at <strong>the</strong> rangewide<br />

scale is coarser-resolution and makes fewer


Figur igur igure igur e 3.1 3.1 Dispersal-distance relationship <strong>for</strong> radio-marked juvenile spotted owls.<br />

Volume II/Chapter 3<br />

FPO<br />

distinctions about habitat details. For example,<br />

many <strong>for</strong>ests have digital elevation models<br />

(DEMs) with a nominal resolution of 30 m (98<br />

ft) and habitat data corresponding to timbermanagement<br />

compartments resolved on <strong>the</strong><br />

order of tens of hectares. At this scale, habitats<br />

may be defined in terms of tree-species composition,<br />

size-class distributions, or o<strong>the</strong>r details. By<br />

contrast, <strong>the</strong> data we have available <strong>for</strong><br />

rangewide analyses have a spatial resolution of 1<br />

km (0.62 miles); habitats at this scale are defined<br />

as gross cover types (e.g., pinyon-juniper). Data<br />

at <strong>the</strong>se two scales lend <strong>the</strong>mselves to <strong>the</strong> same<br />

analytic techniques but require ra<strong>the</strong>r different<br />

interpretation because of <strong>the</strong>ir different resolution<br />

and in<strong>for</strong>mation content.<br />

Data Data Availability<br />

Availability<br />

In keeping with <strong>the</strong> <strong>for</strong>egoing discussion of<br />

scale and resolution, we have attempted to<br />

acquire data at two scales and resolution <strong>for</strong><br />

landscape analyses. At a finer scale, we sought<br />

USGS DEMs with 30-m (98 ft) resolution and<br />

4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

habitat maps of similar spatial resolution and<br />

comparatively high in<strong>for</strong>mation content (e.g.,<br />

species composition, size-class distributions). We<br />

were largely unsuccessful in this ef<strong>for</strong>t and, thus,<br />

have been unable to address questions on spatial<br />

aspects of habitat use by owls at this scale.<br />

At a coarser scale, we acquired two sets of<br />

habitat data. Both sets were derived from<br />

AVHRR satellite imagery and have a spatial<br />

resolution of 1 km 2 (0.39 miles 2 ). One set is <strong>the</strong><br />

EROS Land Cover classification (Loveland et al.<br />

1991), which recognizes 159 cover classes across<br />

<strong>the</strong> conterminous United States. The second<br />

dataset is derived from <strong>the</strong> EROS set but was<br />

reclassified by <strong>the</strong> FS into a smaller number of<br />

recognized Forest Cover Types (Powell et al.<br />

1993, Zhu and Evans 1992, Evans and Zhu<br />

1993). This latter set also includes a companion<br />

coverage of <strong>for</strong>est density (percent canopy cover)<br />

derived by resampling <strong>the</strong> AVHRR-based data<br />

with higher-resolution Thematic Mapper (TM)<br />

imagery at 30-m (98 ft) resolution and correlating<br />

<strong>the</strong> percent of cells <strong>for</strong>ested at 30-m (98 ft)<br />

resolution to a greenness index (NDVI) at 1-km


esolution (Zhu 1994). The FS has also created a<br />

preliminary classification of Mexico’s <strong>for</strong>est cover<br />

types, based on <strong>the</strong> same AVHRR imagery<br />

(Evans et al. 1992). No <strong>for</strong>est density coverage<br />

exists <strong>for</strong> Mexico at this time. Examination of<br />

<strong>the</strong>se coverages revealed that <strong>the</strong> EROS classification<br />

included too much detail and some<br />

details that were ra<strong>the</strong>r suspect from a biogeographic<br />

standpoint, both of which argued<br />

against its use. The FS coverages span <strong>the</strong> entire<br />

range of <strong>the</strong> owl within <strong>the</strong> United States and at<br />

an appropriate in<strong>for</strong>mation content (number of<br />

classes). For <strong>the</strong>se reasons, we have elected to<br />

base our rangewide analyses on <strong>the</strong> FS coverages.<br />

At this scale, we also have 1-km 2 (0.39 miles 2 )<br />

resolution DEMs as provided by EROS. Because<br />

we lack comparable data <strong>for</strong> Mexico, our analyses<br />

are restricted to <strong>the</strong> United States.<br />

The lack of data at higher resolution (e.g.,<br />

30-m scale) has important implications in our<br />

analyses, in that it precludes analysis of those<br />

features too fine-grained to be resolved in <strong>the</strong><br />

coarse-resolution data we have been <strong>for</strong>ced to<br />

use. For example, <strong>the</strong> 1-km scale data cannot<br />

resolve small canyons that are important to owls<br />

in parts of Utah. Nei<strong>the</strong>r do we have data that<br />

can resolve details about owl microhabitat, nor<br />

even subtle distinctions among <strong>for</strong>est cover<br />

types. As we note below, this lack of data does<br />

not render our analyses pointless, but it does<br />

emphasize <strong>the</strong> need to repeat and corroborate<br />

<strong>the</strong>se analyses with finer-resolution data. On <strong>the</strong><br />

o<strong>the</strong>r hand, our coarse-resolution approach<br />

allows us to analyze landscape pattern over<br />

virtually all of <strong>the</strong> owl’s range within <strong>the</strong> United<br />

States.<br />

Landscape Landscape Connectedness<br />

Connectedness<br />

Connectedness<br />

Our approach derives from graph <strong>the</strong>ory and<br />

percolation <strong>the</strong>ory (Gardner et al. 1992) and<br />

focuses on <strong>the</strong> so-called “radius of gyration” of<br />

<strong>the</strong> largest subgraph in a graph representing a<br />

landscape of habitat patches. In this, a “graph” is<br />

a set (map) of habitat clusters, and each “cluster”<br />

is a collection of grid cells that are defined to be<br />

functionally connected. In practice, we define<br />

this connection based on a “minimum joining<br />

distance” such that cells that are within this<br />

distance are functionally connected. The radius<br />

Volume II/Chapter 3<br />

5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

of gyration is defined as <strong>the</strong> mean Euclidean<br />

distance between each cell of a habitat cluster<br />

and <strong>the</strong> centroid of that cluster. Habitat clusters<br />

are <strong>for</strong>med by specifying a joining distance and<br />

<strong>the</strong>n identifying groups of cells that are spatially<br />

discrete at that distance scale. As one increases<br />

this joining distance, <strong>the</strong> habitat map coalesces<br />

into increasingly larger clusters as isolated<br />

patches are subsumed into nearby clusters. A<br />

convenient means of indexing this process is as<br />

<strong>the</strong> radius of <strong>the</strong> largest cluster in <strong>the</strong> graph. A<br />

weighted index <strong>for</strong> <strong>the</strong> map can be computed by<br />

summing <strong>the</strong> radii of all clusters in <strong>the</strong> map,<br />

weighting each radius by <strong>the</strong> proportion of <strong>the</strong><br />

entire map it comprises (i.e., its relative size).<br />

This weighted sum is <strong>the</strong> map’s “correlation<br />

length” (or connectedness length) and has as<br />

units <strong>the</strong> distance units of <strong>the</strong> original map. One<br />

interprets correlation length in terms of how far,<br />

on average, an animal could traverse <strong>the</strong> map<br />

without straying off “habitat” cells into<br />

nonhabitat.<br />

The ultimate goal here is not to compute<br />

landscape connectedness in itself, but ra<strong>the</strong>r to<br />

identify those patches (clusters) that contribute<br />

most significantly to overall habitat connectedness.<br />

One means to this end is to estimate <strong>the</strong><br />

reduction in connectedness that would result if a<br />

patch were removed from <strong>the</strong> landscape. Here,<br />

this estimate is computed as <strong>the</strong> reduction in<br />

correlation length of <strong>the</strong> landscape. These<br />

reductions are estimated by systematically<br />

removing each cluster and recomputing <strong>the</strong><br />

connectedness index. The analysis itself proceeds<br />

in steps as follows:<br />

1. Define a binary raster map of “habitat”<br />

versus “nonhabitat.”<br />

2. Specify a distance threshold at which<br />

patches may join.<br />

3. Per<strong>for</strong>m <strong>the</strong> cluster-removal experiment,<br />

saving each cluster’s effect on <strong>the</strong> correlation<br />

length.<br />

4. Normalize this effect <strong>for</strong> each cluster by<br />

dividing its effect by its total area, and<br />

save this value as well.


The normalization in step (4) adjusts <strong>the</strong><br />

patch’s effect <strong>for</strong> its area, which identifies those<br />

patches whose effect on connectedness is large<br />

relative to <strong>the</strong>ir size. The result of <strong>the</strong> analysis is<br />

a ranking of habitat patches in terms of <strong>the</strong><br />

reduction in connectedness each elicits, that is,<br />

each patch’s contribution to overall connectedness.<br />

This procedure is <strong>the</strong>n repeated <strong>for</strong> a range<br />

of threshold distances to estimate <strong>the</strong> consequences<br />

of varying assumptions about <strong>the</strong><br />

dispersal capabilities of owls. We used distances<br />

from 0 to 100 km (0-62 miles) in 5-km (3.1<br />

mile) intervals. This range spans our best empirical<br />

estimate of <strong>the</strong> owl’s dispersal range, which is<br />

on <strong>the</strong> order of 50 km (31 miles). If <strong>the</strong> patch<br />

ranks change considerably at different joining<br />

distances, this would suggest a need to improve<br />

<strong>the</strong> accuracy of our estimates of owl dispersal<br />

distances so that we could tailor <strong>the</strong> analysis to<br />

maximize its biological significance.<br />

Finally, <strong>the</strong> entire analysis is repeated <strong>for</strong> a<br />

variety of baseline habitat maps. This gives us an<br />

estimate of how robust <strong>the</strong> results are to assumptions<br />

about what constitutes “potential owl<br />

habitat.” At this spatial scale and with <strong>the</strong> data<br />

available, we have limited opportunities to devise<br />

alternative definitions of “owl habitat.” We have<br />

devised three alternative habitat maps:<br />

1. A map wherein all cells assigned as<br />

“Douglas-fir” (i.e., mixed conifer) or<br />

ponderosa pine (including some mixedconifer<br />

as well as pine-oak) cover types<br />

are defined to be potential owl habitat;<br />

2. A map including all Douglas-fir and<br />

ponderosa pine cells, plus any pinyonjuniper<br />

cells that have greater than 50%<br />

canopy cover as estimated in <strong>the</strong> FS<br />

Forest Density coverage;<br />

3. A map including all Douglas-fir and<br />

ponderosa pine types, plus those pinyonjuniper<br />

cells that are within a 2-km (1.2<br />

miles) buffer of Douglas-fir or pine<br />

types, i.e., pinyon-juniper near better<br />

owl habitat.<br />

Volume II/Chapter 3<br />

6<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

These alternative definitions vary in terms of<br />

how narrowly or generously “habitat” is defined.<br />

O<strong>the</strong>r definitions could certainly be devised, and<br />

we do not argue that ours are <strong>the</strong> only (nor even<br />

<strong>the</strong> best) possibilities. We will argue, however,<br />

that <strong>the</strong>se are sufficient to indicate whe<strong>the</strong>r our<br />

conclusions are robust to <strong>the</strong> definition of<br />

habitat or whe<strong>the</strong>r we need to invest more ef<strong>for</strong>t<br />

in generating more realistic base habitat maps.<br />

Results Results of of Landscape Landscape Analyses<br />

Analyses<br />

Results of <strong>the</strong>se analyses provide insight into<br />

two keys aspects of habitat connectedness: (1)<br />

<strong>the</strong> relationship between minimum joining<br />

distance and overall connectedness as indexed by<br />

correlation length; and (2) contributions of<br />

individual habitat patches to overall connectedness.<br />

We present each of <strong>the</strong>se in turn. Because<br />

<strong>the</strong> results were qualitatively similar <strong>for</strong> each of<br />

<strong>the</strong> baseline habitat maps, we present here only<br />

<strong>the</strong> results <strong>for</strong> <strong>the</strong> mixed-conifer/ponderosa pine<br />

base map; but we return to this issue later in our<br />

discussion.<br />

Landscape Landscape Connectedness Connectedness and and and Minimum<br />

Minimum<br />

Joining Joining Distance<br />

Distance<br />

Correlation length exhibits a profoundly<br />

nonlinear relationship with minimum joining<br />

distance in all cases we examined (Figure 3.2).<br />

The inflection in this relationship illustrates that<br />

<strong>the</strong> landscape changes from being largely “unconnected”<br />

to largely “connected” over a narrow<br />

range of distances. In <strong>the</strong> habitat maps we<br />

analyzed, this transition occurred over distances<br />

of 40-60 km (25-37 miles). This result varied<br />

only slightly <strong>for</strong> alternative habitat maps, suggesting<br />

that this qualitative result is ra<strong>the</strong>r robust<br />

to <strong>the</strong>se definitions.<br />

The relationship between cluster size and<br />

joining distance varies systematically with <strong>the</strong><br />

spatial resolution with which habitats are defined.<br />

This can be seen most clearly by contrasting<br />

<strong>the</strong> curve <strong>for</strong> “all clusters” of ponderosa pine<br />

and Douglas-fir as compared to <strong>the</strong> curve <strong>for</strong> <strong>the</strong><br />

largest 254 of <strong>the</strong>se clusters (Figure 3.2). Including<br />

more (smaller) clusters shifts <strong>the</strong> curve to <strong>the</strong><br />

left, effectively joining <strong>the</strong> landscape at closer


distances. Presumably, if we were to include very<br />

small clusters <strong>the</strong> landscape might be functionally<br />

connected at extremely small joining distances,<br />

which is something of a scaling artifact.<br />

This result does underscore <strong>the</strong> need to more<br />

fully characterize <strong>the</strong> habitat affinities and<br />

dispersal behavior of owls. If we could limit<br />

cluster sizes and joining (dispersal) distances to<br />

biologically reasonable values, this analysis<br />

would be more confidently focused.<br />

All habitat maps coalesced into a very few<br />

large clusters at joining distances of approximately<br />

60 km (37 miles) or more. This indicates<br />

that at <strong>the</strong>se distances, <strong>the</strong> entire landscape is<br />

essentially connected (Figure 3.3). Yet even at<br />

joining distances of 100 km (62 miles), a few<br />

clusters remain spatially discrete, and by implication,<br />

functionally isolated from <strong>the</strong> rest of <strong>the</strong><br />

habitat in <strong>the</strong> landscape.<br />

Rank Rank Rank Patch Patch Contributions Contributions to to Connectedness<br />

Connectedness<br />

Connectedness<br />

The influence of each patch on overall<br />

connectedness was tallied <strong>for</strong> only <strong>the</strong> largest<br />

254 clusters in each map. In fact, <strong>the</strong> maps<br />

contained thousands of clusters but many were<br />

single cells; so it proved to be computationally<br />

impractical to compute every case. Patch influence<br />

on correlation length was strongly dependent<br />

on joining distance: <strong>the</strong> average influence<br />

was greatest at intermediate distance scales and<br />

much less at shorter or longer scales (Figure 3.4).<br />

To highlight influential patches spatially, we<br />

produced a habitat map in which each patch is<br />

color-coded according to its mean rank over all<br />

distance classes (Figure 3.5a). A similar map<br />

emphasizes patch importance to connectedness<br />

at <strong>the</strong> intermediate distance scale of 45 km (28<br />

miles) (Figure 3.5b). In both maps, “hot” colors<br />

indicate highly ranked patches (those contributing<br />

substantially to landscape connectedness)<br />

while “cool” colors indicate patches of low rank<br />

(those with little effect on connectedness).<br />

In general, patch effects on connectedness<br />

depended on patch area, in that large patches<br />

had <strong>the</strong> greatest influence on overall connectedness.<br />

Patch ranks, normalized <strong>for</strong> area, are<br />

illustrated as averaged over all distances (fig.<br />

3.6a) and at 45-km (28-mile) joining distance<br />

(Figure 3.6b), using <strong>the</strong> same color scheme as in<br />

Volume II/Chapter 3<br />

7<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure 3.5. These figures emphasize <strong>the</strong> importance<br />

of a few clusters in joining <strong>the</strong> large habitat<br />

block along <strong>the</strong> Mogollon Rim to <strong>the</strong> large<br />

cluster far<strong>the</strong>r nor<strong>the</strong>ast.<br />

DISCUSSION<br />

DISCUSSION<br />

Distance Distance Relationships Relationships in in in Landscape<br />

Landscape<br />

Connectedness<br />

Connectedness<br />

The nonlinear relationship in Figure 3.2<br />

makes intuitive sense if one envisions <strong>the</strong> process<br />

that generates this relationship. At small joining<br />

distances, small and nearby patches are joined<br />

into somewhat larger clusters, but <strong>the</strong> landscape<br />

still consists mostly of disjoint clusters. At a<br />

particular distance, a large subset of <strong>the</strong> clusters<br />

joins into a single large cluster. Once this has<br />

happened, fur<strong>the</strong>r small accretions to <strong>the</strong> cluster<br />

do not change its total area appreciably. From a<br />

functional standpoint, <strong>the</strong>se later additions are<br />

redundant links to an “already connected”<br />

cluster. This same process explains why <strong>the</strong><br />

graph of average influence of patch removal<br />

(Figure 3.4) shows a peak at <strong>the</strong>se intermediate<br />

distance classes; <strong>for</strong> shorter or longer distances<br />

<strong>the</strong>se patches can have little impact on overall<br />

connectedness.<br />

An important result of this analysis is that<br />

<strong>the</strong> strongly nonlinear domain of this relationship<br />

(i.e., <strong>the</strong> inflection point in Figure 3.2)<br />

coincides with our best current estimate of <strong>the</strong><br />

dispersal capabilities of spotted owls based on<br />

field studies, approximately 50 km (31 miles).<br />

By implication, if owls’ dispersal ranges were<br />

much less than this estimate (say, 20 km [12<br />

miles]) <strong>the</strong>n much of <strong>the</strong>ir natural range would<br />

be functionally unconnected (most patches<br />

would be isolated). Reciprocally, if owls could<br />

disperse much far<strong>the</strong>r than our current estimate<br />

(say, 100 km [62 miles]), <strong>the</strong>n most of <strong>the</strong><br />

landscape would be functionally connected.<br />

Likewise, <strong>the</strong> change in this relationship with <strong>the</strong><br />

inclusion of more, smaller patches suggests that<br />

owls’ use of very small patches as dispersal<br />

conduits or stepping stones could influence <strong>the</strong>se<br />

results. If owls can use very small patches, <strong>the</strong>n<br />

<strong>the</strong> landscape might be more connected than our


Volume II/Chapter 3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 3.2. 3.2. The relationship between correlation length and minimum joining distance. The 4 lines<br />

illustrate <strong>the</strong> efects of alternative definitions of “potential owl habitat.”<br />

FPO<br />

results suggest, presuming, of course, that such<br />

small patches exist.<br />

The radio-telemetry data suggest that juvenile<br />

owls have <strong>the</strong> dispersal capability to provide<br />

population and genetic links between subpopulations.<br />

To date, none of <strong>the</strong> radio-marked<br />

juveniles has been recruited into a resident,<br />

territorial subpopulation. In attempting to judge<br />

whe<strong>the</strong>r owl dispersal is sufficient to maintain<br />

population and genetic connectedness, we are<br />

faced with <strong>the</strong> logistical problem that ecologically<br />

significant dispersal events might occur<br />

quite infrequently (1-2 per generation) and<br />

would likely go unrecorded by even <strong>the</strong> most<br />

intensive monitoring program. Thus, our data<br />

can suggest that such dispersal capabilities might<br />

exist, but <strong>the</strong>se data cannot prove that such<br />

dispersal actually occurs. Importantly, nei<strong>the</strong>r<br />

can our lack of data prove that such dispersal<br />

does not occur.<br />

Clearly <strong>the</strong>se results point to a need <strong>for</strong><br />

better understanding of <strong>the</strong> dispersal behavior<br />

and distance relationships <strong>for</strong> spotted owls. One<br />

source of uncertainty in our analyses is that <strong>the</strong><br />

8<br />

clustering is based on boolean distance decisions<br />

(i.e., patches are connected if <strong>the</strong>ir distance is<br />

strictly within <strong>the</strong> joining distance). But animal<br />

dispersal is probabilistic, exhibiting some sort of<br />

decreasing probability with increasing distance<br />

(recall Figure 3.1). This distinction exaggerates<br />

our results relative to how dispersal probably<br />

occurs with real animals.<br />

A second source of uncertainty stems from<br />

our limited understanding of owl dispersal<br />

behavior. Our analyses are based on assumptions<br />

about “reachability” with <strong>the</strong> tacit assumption<br />

being that if habitat is reachable in terms of<br />

absolute distance, <strong>the</strong>n owls can and will disperse<br />

<strong>the</strong>re. But <strong>the</strong>re are many plausible reasons why<br />

this might not be so (e.g., avoidance behavior,<br />

excessive mortality during dispersal); so it<br />

remains that we need to temper our interpretations<br />

with additional considerations of owl<br />

dispersal.


Volume II/Chapter 3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 3.3. 3.3. Landscape mosaics of discrete clusters of Douglas-fir and Ponderosa pine habitat types, as<br />

(a) (a) largest 254 clusters, and at (b) (b) 30-km, and (c) (c) 60-km joining distances. Colors have no significance<br />

beyond labelling discrete clusters.<br />

FPO<br />

9


Volume II/Chapter 3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 3.4. 3.4. Change in correlation length due to cluster removal as a function of distance, averaged<br />

over <strong>the</strong> largest 254 clusters.<br />

FPO<br />

Patch Patch Contributions Contributions to to Landscape<br />

Landscape<br />

Connectedness<br />

Connectedness<br />

Rank scores <strong>for</strong> patch contributions to<br />

overall landscape connectedness make intuitive<br />

sense when viewed as landscape mosaics (Figures<br />

3.5 and 3.6). The ranks uncorrected <strong>for</strong> area<br />

show <strong>the</strong> expected result that large clusters that<br />

are centrally located have <strong>the</strong> highest ranks,<br />

while small or isolated clusters are lower-ranked.<br />

Area-normalized ranks emphasize <strong>the</strong> positional<br />

aspects of patch contributions and Figure 3.6<br />

illustrates a few patches (in red) that seem to act<br />

as bridges spanning larger habitat areas. Some of<br />

<strong>the</strong>se bridges are ra<strong>the</strong>r small yet could be<br />

important links in landscape connectedness.<br />

An important caveat to bear in mind is that<br />

<strong>the</strong>se analyses are all based on habitat, not on<br />

owl densities. Thus, sparsely populated habitat<br />

clusters in <strong>the</strong> nor<strong>the</strong>rn reaches of <strong>the</strong> owl’s<br />

range receive equal weight in <strong>the</strong> analysis as<br />

compared to habitats currently supporting much<br />

larger owl populations, <strong>for</strong> example, along <strong>the</strong><br />

10<br />

Mogollon Rim. Thus, <strong>the</strong> clusters in Colorado<br />

that appear important to connectedness (red<br />

patches in <strong>the</strong> upper-right regions of Figure 3.5)<br />

are actually connecting habitat that supports<br />

essentially no owls. If habitat clusters were<br />

weighted according to present-day owl abundance,<br />

<strong>the</strong>se same analyses would provide quite<br />

different results. Densely populated patches<br />

would increase in importance while more<br />

sparsely populated patches would be downweighted.<br />

While this is ra<strong>the</strong>r easy to anticipate<br />

in general, such a weighted analysis would<br />

require much better spatial in<strong>for</strong>mation on owl<br />

abundance across its range than <strong>the</strong> inconsistent<br />

coverage currently available.<br />

These considerations bear strongly on <strong>the</strong><br />

ultimate goal of a conservation plan. If our goal<br />

is to provide a template that might sustain owl<br />

metapopulations well into <strong>the</strong> future, <strong>the</strong>n an<br />

analysis weighted on “potential habitat” seems<br />

most appropriate. Conversely, a strategy to<br />

preserve current populations would seem to<br />

argue <strong>for</strong> an analysis heavily weighted on<br />

present-day owl abundance. Our results suggest


Volume II/Chapter 3<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 3.5. 3.5. Maps with clusters colored to illustrate <strong>the</strong>ir rank importance, weighted by (uncorrected<br />

<strong>for</strong>) patch area. Hot colors (reds) are highly ranked; cool colors (blue-violet), low ranked. Base habitat<br />

map is ponderosa pine/Douglas fir. (a) (a) Patch importance averaged of all distance classes. (b) (b) Importance<br />

at 45-km joining distance.<br />

11


Volume II/Chapter 3<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 3.6. 3.6. Patch importance to overall connectedness, normalized <strong>for</strong> patch area. Color scheme and<br />

base map, and panels (a) (a) and (b) (b) (b), (b) are <strong>the</strong> same as in Figure 3.5.<br />

12


that <strong>the</strong>se two strategies might lead to quite<br />

different recommendations.<br />

Importantly, our rankings of patch importance<br />

to overall connectedness are based on a<br />

simple patch-removal algorithm in which each<br />

patch is removed singly from <strong>the</strong> existing landscape.<br />

Clearly, <strong>the</strong>se results could be quite<br />

different if <strong>the</strong> algorithm provided <strong>for</strong> more<br />

complex scenarios as conditional removals. For<br />

example, if patch i has already been removed,<br />

<strong>the</strong>n <strong>the</strong> removal of patch j might take on much<br />

greater importance. Such conditional scenarios<br />

would be more realistic in <strong>the</strong> sense that landscape<br />

dynamics driven by land-use management<br />

are time-structured (sequential), conditional, and<br />

typically act on multiple patches during any<br />

single management episode. Such complicated<br />

scenarios might be undertaken on a smaller scale<br />

(e.g., <strong>for</strong> a National Forest) if sufficient data were<br />

available <strong>for</strong> <strong>the</strong> analysis. Complicated, conditional<br />

scenarios are probably not feasible <strong>for</strong><br />

rangewide analyses.<br />

Sensitivity Sensitivity to to Habitat Habitat Definition<br />

Definition<br />

Two empirical biases emerge in considering<br />

alternative definitions of what constitutes “potential<br />

owl habitat” in <strong>the</strong>se analyses. One bias is<br />

due to <strong>the</strong> spatial scale at which habitat patches<br />

are resolved. As smaller patches are included,<br />

overall landscape connectedness tends to increase<br />

so long as <strong>the</strong>se small patches are liberally<br />

sprinkled across <strong>the</strong> landscape. Similarly, <strong>for</strong><br />

habitat definitions that are increasingly “generous”<br />

toward owls, more potential owl habitat<br />

occurs in <strong>the</strong> landscape and so connectedness<br />

also tends to increase (consider a map that<br />

includes some pinyon-juniper relative to <strong>the</strong><br />

mixed-conifer/pine landscape). These biases<br />

appear in our analyses as a result of data availability.<br />

If higher-resolution data were available,<br />

more patches could possibly be delineated. At<br />

<strong>the</strong> same time, however, given higher-resolution<br />

data we could define owl habitat more stringently<br />

and thus some patches would be redefined<br />

as no longer usable by owls; owl habitat<br />

would decrease in abundance and overall connectedness<br />

would decrease accordingly. Thus, <strong>the</strong><br />

two empirical biases are somewhat compensat-<br />

Volume II/Chapter 3<br />

13<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

ing. But both biases point to a need to improve<br />

our ability to discriminate usable owl habitat<br />

from <strong>the</strong> surrounding matrix.<br />

O<strong>the</strong>r O<strong>the</strong>r Uncertainties Uncertainties and<br />

and<br />

Considerations<br />

Considerations<br />

A final consideration of our results must<br />

address any uncertainties or biases that might<br />

result from <strong>the</strong> algorithm itself. One potential<br />

bias that emerges can be seen in <strong>the</strong> figures<br />

illustrating patch importance to connectedness.<br />

Because our approach indexes connectedness as<br />

<strong>the</strong> mean size of <strong>the</strong> largest cluster, a bias<br />

emerges whereby patches that are peripheral in<br />

<strong>the</strong> landscape can <strong>for</strong>m clusters with a very large<br />

radius. Thus, <strong>the</strong> important patches in Figure<br />

3.5 tend to <strong>for</strong>m a ring around <strong>the</strong> landscape as a<br />

whole, partly because <strong>the</strong>se patches are large but<br />

also because <strong>the</strong>ir joining creates a cluster that<br />

has a radius nearly as large as <strong>the</strong> entire mosaic.<br />

This bias would not occur if <strong>the</strong> index of connectedness<br />

counted total area in a way that did<br />

not emphasize among-cell distances. For example,<br />

an index that estimated “total connected<br />

area” ra<strong>the</strong>r than <strong>the</strong> effective size of this area<br />

might yield a slightly different estimate of patch<br />

importance. Un<strong>for</strong>tunately, such indices have<br />

proven to be computationally unfeasible thus far.<br />

We continue to explore alternative algorithms<br />

<strong>for</strong> indexing habitat connectedness.<br />

CONCLUSIONS<br />

CONCLUSIONS<br />

Application of metapopulation <strong>the</strong>ory to <strong>the</strong><br />

<strong>Mexican</strong> spotted owl is ra<strong>the</strong>r speculative, given<br />

our limited understanding of within-population<br />

dynamics much less between-population dynamics.<br />

If metapopulation models actually represent<br />

realistic abstractions of real-world processes, <strong>the</strong>n<br />

a number of different metapopulation models<br />

may apply to different geographic regions within<br />

<strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong> spotted owl. For<br />

example, one could envision a core-satellite<br />

model applying to <strong>the</strong> sou<strong>the</strong>rn portion of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl range with <strong>the</strong> Upper Gila<br />

RU acting as a core source population with <strong>the</strong><br />

smaller surrounding mountain ranges in Basin


and Range - West and Colorado Plateau RUs<br />

acting as satellite sinks. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong><br />

Sky Island mountain ranges in sou<strong>the</strong>rn Arizona<br />

could also follow <strong>the</strong> classic metapopulation<br />

dynamics proposed by Levins (1969). A number<br />

of different scenarios could be envisioned,<br />

which, un<strong>for</strong>tunately, we cannot corroborate<br />

easily empirically. The distribution of geographically<br />

isolated subpopulations, however, suggests<br />

that some <strong>for</strong>m of interaction between <strong>the</strong>se<br />

subpopulations is plausible. Clearly, intensive,<br />

long-term studies over large areas will be required<br />

to understand <strong>the</strong> structure and dynamics<br />

of <strong>Mexican</strong> spotted owl population at <strong>the</strong>se large<br />

scales.<br />

Although <strong>the</strong> landscape analyses are exploratory,<br />

some general conclusions can still be<br />

drawn from <strong>the</strong> results. These conclusions<br />

concern apparent connectedness of various<br />

regions of <strong>the</strong> owl’s range and <strong>the</strong> relative importance<br />

to particular habitat clusters to overall<br />

landscape connectedness.<br />

Regardless of <strong>the</strong> underlying habitat map<br />

used in analyses, results consistently show a few<br />

regions that appear functionally isolated at<br />

intermediate joining distances similar to <strong>the</strong><br />

dispersal range of owls. For example, large blocks<br />

of sou<strong>the</strong>rn Utah persist as discrete habitat<br />

clusters at joining distances of 40 km (25 miles);<br />

<strong>the</strong> Lincoln National Forest also appears isolated<br />

at this spatial scale. We might test <strong>the</strong> hypo<strong>the</strong>sis<br />

that <strong>the</strong>se subpopulations are discrete, possibly<br />

by exploring genetic similarities between <strong>the</strong>se<br />

and more central (connected) populations such<br />

as those along <strong>the</strong> Mogollon Rim. Likewise,<br />

basic population analyses of <strong>the</strong>se populations<br />

might also indicate <strong>the</strong>ir degree of functional<br />

connectedness. For example, spatial<br />

discontinuities in population density, age structure,<br />

or o<strong>the</strong>r parameters might suggest that<br />

<strong>the</strong>se populations do not interact to <strong>the</strong> same<br />

extent as o<strong>the</strong>r, more contiguous populations.<br />

The relative importance of individual habitat<br />

patches, when uncorrected <strong>for</strong> patch area,<br />

suggests <strong>the</strong> intuitive approach of protecting<br />

those patches that currently support <strong>the</strong> highest<br />

owl densities, such as along <strong>the</strong> Mogollon Rim.<br />

But our results also indicate a high importance<br />

<strong>for</strong> patches far<strong>the</strong>r north in <strong>the</strong> owl’s range,<br />

Volume II/Chapter 3<br />

14<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

patches which currently do not support appreciable<br />

owl populations. Our reaction to this<br />

result depends in part on whe<strong>the</strong>r our goal is to<br />

preserve present-day populations or to protect<br />

<strong>the</strong> capability <strong>for</strong> populations to expand in <strong>the</strong><br />

future. The discrepancy between <strong>the</strong>se two<br />

strategies is most pronounced in <strong>the</strong> nor<strong>the</strong>rn<br />

reaches of <strong>the</strong> owl’s current range.<br />

Correcting cluster importance <strong>for</strong> area<br />

indicates <strong>the</strong> contribution, per unit area, of each<br />

of <strong>the</strong> habitat clusters. This correction suggests<br />

that while <strong>the</strong> cluster along <strong>the</strong> Mogollon Rim is<br />

crucial to overall connectedness, <strong>the</strong> many stands<br />

making up this cluster are not so important on<br />

an individual basis. Thus, this cluster would<br />

likely continue to play an important role in <strong>the</strong><br />

landscape so long as its internal continuity is<br />

maintained, which would require reiterating our<br />

analyses on a finer spatial scale and resolution as<br />

management of this region proceeds. Conversely,<br />

a few clusters emerge as being more important<br />

per unit area than <strong>the</strong>ir uncorrected importance<br />

would suggest, such as <strong>the</strong> stepping-stones<br />

evident in Figure 3.6 as red patches. Such<br />

patches warrant special attention in land use<br />

planning because of <strong>the</strong>ir potential to affect<br />

regional populations despite <strong>the</strong>ir small size and<br />

perhaps modest owl populations. Especially,<br />

<strong>the</strong>se patches should be a focus of monitoring<br />

ef<strong>for</strong>ts so that <strong>the</strong>ir use by owls can be assessed.<br />

Note that <strong>the</strong>se patches typically would not be<br />

targeted in field studies <strong>for</strong> <strong>the</strong> simple reason<br />

that <strong>the</strong>y would not appear to support large owl<br />

populations.<br />

Finally, we should emphasize that <strong>the</strong>se<br />

results are exploratory and <strong>the</strong>re<strong>for</strong>e subject to<br />

corroboration. One <strong>for</strong>m of verification could<br />

come via analyses of owl subpopulations across<br />

<strong>the</strong> region. Metapopulation <strong>the</strong>ory suggests that<br />

isolated habitats should show higher year-to-year<br />

variability in population density than would<br />

better-connected patches, which would be better<br />

subsidized by dispersal. A second <strong>for</strong>m of corroboration<br />

would entail analyses similar to ours<br />

but using alternative indices of connectedness<br />

and alternative definitions of suitable owl habitat.<br />

Especially, we would like to see <strong>the</strong>se analyses<br />

repeated with higher-resolution habitat data<br />

capable of resolving <strong>the</strong> those features most<br />

important to owls. In any case, it is clear that we


need to invest much more ef<strong>for</strong>t toward defining<br />

<strong>the</strong> dispersal capabilities and behavior of <strong>the</strong><br />

spotted owl across its range.<br />

In closing, we would like to revisit <strong>the</strong><br />

rationale that underlies our somewhat <strong>the</strong>oretical,<br />

habitat-based approach. While it might be<br />

appealing to invoke analyses that rely on owl<br />

population data, we were <strong>for</strong>ced to accept at <strong>the</strong><br />

beginning that we lacked adequate data <strong>for</strong> such<br />

an analysis across <strong>the</strong> owl’s range. Similarly,<br />

detailed population models such as those developed<br />

<strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn owl are appealing because<br />

of <strong>the</strong>ir biological richness; but we clearly lack<br />

<strong>the</strong> data to parameterize such a model <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. We see both <strong>the</strong>se population-based<br />

approaches as a useful adjunct to our<br />

approach, but one that we cannot support<br />

empirically with data currently available. Instead,<br />

we developed an approach that makes very few<br />

assumptions about owl biology, and we tested<br />

our results to determine whe<strong>the</strong>r violation of<br />

<strong>the</strong>se assumptions would change our results<br />

substantially. Our approach was to use as generous<br />

a definition of owl habitat as possible given<br />

our data; our major conclusions seem largely<br />

unaffected by alternative definitions of habitat,<br />

although this needs to be verified with finerresolation<br />

data. The o<strong>the</strong>r important assumption<br />

we made was that owls have a dispersal-distance<br />

relationship such that dispersal probability<br />

decreases with increasing distance, and that <strong>the</strong>ir<br />

average dispersal is in <strong>the</strong> neighborhood of 50<br />

km or so (i.e., not > 100<br />

km). Within this domain, our results are also<br />

robust. Thus, while <strong>the</strong>re remain a number of<br />

questions still to be resolved concerning landscape-scale<br />

patterns and owl metapopulations,<br />

<strong>the</strong> results we present here are a useful and valid<br />

first approximation.<br />

Volume II/Chapter 3<br />

15<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

LITERATURE LITERATURE CITED<br />

CITED<br />

Doak, D. F., and L. S. Mills. 1994. A useful role<br />

<strong>for</strong> <strong>the</strong>ory in conservation. Ecology<br />

75:615-626.<br />

Evans, D. L., and Z. Zhu. 1993. AVHRR <strong>for</strong><br />

<strong>for</strong>est mapping: national applications<br />

and global implications. Pages 76-79 in<br />

A. Lewis, ed. Looking to <strong>the</strong> future with<br />

an eye on <strong>the</strong> past: proceedings of <strong>the</strong><br />

1993 ACMS/ASPRS. ASPRS.<br />

——., Z. Zhu, S. Eggen-McIntosh, P. G. Mayoral,<br />

and J. L. Omelas de Anda. 1992.<br />

Mapping Mexico’s <strong>for</strong>est lands with<br />

advanced very high resolution radiometer.<br />

USDA For. Serv. Sou<strong>the</strong>rn For. Exp.<br />

Stn., New Orleans, La.<br />

Gardner, R. H., M. G. Turner, V. H. Dale, and<br />

R. V. O’Neill. 1992. A percolation<br />

model of ecological flows. Pages 259-269<br />

in A. J. Hansen and F. di Castri, eds.<br />

Landscape boundaries. Springer-Verlag,<br />

New York, N.Y.<br />

Hanski, I., and M. Gilpin. 1991.<br />

Metapopulation dynamics: brief<br />

history and conceptual domain. Biol. J.<br />

Linnean Soc. 42:3-16.<br />

Harrison, S. 1991. Local extinction in a<br />

metapopulation context: an empirical<br />

evaluation. Biol. J. Linnean Soc. 42:73-<br />

88.<br />

Hodgson, A., and P. Stacey. 1994. Habitat use<br />

and dispersal by <strong>Mexican</strong> spotted owls.<br />

U.S. For. Serv., Rocky Mtn. For. Rng.<br />

Exp. Stn., Flagstaff, Ariz.<br />

Levins, R. 1969. Some demographic and genetic<br />

consequences of environmental heterogeneity<br />

<strong>for</strong> biological control. Bull.<br />

Entomol. Soc. Amer. 15:237-240.<br />

Loveland, T. R., J. W. Merchant, D. O. Ohlen,<br />

and J. F. Brown. 1991. Development of a<br />

land-cover characteristics database <strong>for</strong> <strong>the</strong><br />

conterminous U.S. Photogrammetric<br />

Eng. and Remote Sensing 57:1453-<br />

1463.<br />

Noon, B. R., and K. S. McKelvey. 1992. Stability<br />

properties of <strong>the</strong> spotted owl<br />

metapopulation in sou<strong>the</strong>rn Cali<strong>for</strong>nia.<br />

Pages 187-206 in J. Verner, K. S.


McKelvey, B. R. Noon, R. J. Gutiérrez, G. I.<br />

Gould, Jr., and T. W. Beck, eds. The<br />

Cali<strong>for</strong>nia spotted owl: a technical<br />

assessment of its current status. USDA<br />

For. Serv. Gen. Tech. Rep. PSW-133,<br />

Pac. Southwest Res. Stat., Albany, Calif.<br />

Powell, D. S., J. L. Faulkner, D. Darr, Z. Zhu,<br />

amd D. W. MacCleery. 1993. Forest<br />

resources of <strong>the</strong> United States. USDA<br />

For. Serv. Gen Tech Rep. RM-234,<br />

Rocky Mtn. For. Range Exp. Stat., Ft.<br />

Collins, Colo.<br />

Pulliam, H. R. 1988. Sources, sinks, and population<br />

regulation. Am. Nat. 132:652-<br />

661.<br />

Reynolds, R. T., and C. L. Johnson. 1994.<br />

Distribution and ecology of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl in Colorado: annual report.<br />

USDA For. Serv., Rocky Mtn. For. Rng.<br />

Exp. Stn., Ft. Collins, Colo.<br />

Schaffer, M. L. 1985. The metapopulation and<br />

species conservation: <strong>the</strong> special case of<br />

<strong>the</strong> nor<strong>the</strong>rn spotted owl. Pages 86-99 in<br />

Volume II/Chapter 3<br />

16<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

R. J. Gutiérrez and A. B. Carey, eds.<br />

Ecology and management of <strong>the</strong> nor<strong>the</strong>rn<br />

spotted owl in <strong>the</strong> Pacific Northwest.<br />

USDA For. Serv. Gen. Tech. Rep. PNW-<br />

185, Portland, Oreg.<br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner.<br />

1990. A conservation strategy <strong>for</strong> <strong>the</strong><br />

spotted owl. U.S. Gov. Print. Off.,<br />

Washington, D.C. 427pp.<br />

Willey, D. W. 1993. Home-range characteristics<br />

and juvenile dispersal ecology of <strong>Mexican</strong><br />

spotted <strong>Owl</strong>s in sou<strong>the</strong>rn Utah.<br />

Unpubl. Rep., Utah Div. Wildl. Resour.,<br />

Salt Lake City.<br />

Zhu, Z. 1994. Forest density mapping in <strong>the</strong><br />

lower 48 states: a regression procedure.<br />

USDA For. Serv. Res. Pap. SO-280,<br />

Sou<strong>the</strong>rn For. Exp. Stat., New Orleans,<br />

Louis.<br />

——., and D. L. Evans. 1992. Mapping<br />

midsouth <strong>for</strong>est distributions with<br />

AVHRR data. J. For. 90:27-30.


Volume II/Chapter 4<br />

CHAPTER CHAPTER 4: 4: Habitat Habitat Relationships Relationships of of <strong>the</strong><br />

<strong>the</strong><br />

<strong>Mexican</strong> <strong>Mexican</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong>: <strong>Owl</strong>: Current Current Knowledge<br />

Knowledge<br />

The <strong>Mexican</strong> spotted owl was listed as<br />

threatened primarily due to concerns over loss of<br />

habitat (USDI 1993). Consequently, understanding<br />

<strong>the</strong> habitat relationships of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl is critical to developing sound<br />

management plans <strong>for</strong> this species. Here, we<br />

summarize both recent and historical in<strong>for</strong>mation<br />

on habitat relationships of <strong>Mexican</strong> spotted<br />

owls. Because most historical in<strong>for</strong>mation on <strong>the</strong><br />

owl’s habitat contains little quantitative in<strong>for</strong>mation,<br />

we rely mainly on recent in<strong>for</strong>mation<br />

(1985 - present time).<br />

Our primary objectives in this treatment are<br />

to: (1) evaluate and describe patterns of habitat<br />

use by <strong>Mexican</strong> spotted owls; (2) evaluate and<br />

describe patterns of habitat selection by <strong>Mexican</strong><br />

spotted owls; (3) identify specific habitats or<br />

habitat components that appear to be particularly<br />

important to <strong>the</strong> owl; and (4) identify areas<br />

where fur<strong>the</strong>r in<strong>for</strong>mation on habitat relationships<br />

of this owl are most needed. We use <strong>the</strong><br />

terms habitat, habitat use, and habitat selection<br />

as defined by Block and Brennan (1993). Thus,<br />

habitat as used here refers to “<strong>the</strong> subset of<br />

physical environmental factors that a species<br />

requires <strong>for</strong> its survival and reproduction” (Block<br />

and Brennan 1993:36). Habitat use refers to “<strong>the</strong><br />

manner in which a species uses a collection of<br />

environmental components to meet life requisites”,<br />

and habitat selection refers to “disproportional<br />

use of environmental conditions” (Block<br />

and Brennan 1993:38).<br />

Patterns of habitat use evaluated here are<br />

largely descriptive. We evaluated patterns of<br />

habitat selection by comparing use of vegetation<br />

types or specific habitat components to <strong>the</strong><br />

occurrence of those types or components.<br />

Ecological patterns and processes vary with<br />

spatial scale (Urban et al. 1987, O’Neill et al.<br />

1988, Turner 1989), however, and considering<br />

patterns of resource use at only one scale can<br />

yield misleading results (Porter and Church<br />

1987, Orians and Wittenberger 1991, Block and<br />

Brennan 1993). There<strong>for</strong>e, we used a number of<br />

data sources, both published and unpublished,<br />

Joseph L. Ganey and James L. Dick, Jr.<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

to examine habitat use and/or selection at five<br />

spatial scales. At some scales we could evaluate<br />

habitat selection, whereas at o<strong>the</strong>rs we could<br />

only describe patterns of habitat use. Scales<br />

examined are described below, arrayed from<br />

coarsest to finest scale:<br />

1. Landscape scale - habitat use across <strong>the</strong><br />

entire range of <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

2. Home-range scale - patterns of habitat<br />

use within owl home ranges as defined<br />

by locations of radio-tagged owls.<br />

3. Stand scale - use of relatively homogeneous<br />

units of <strong>for</strong>est vegetation within<br />

owl home ranges.<br />

4. Site scale - habitat use proximate to nest,<br />

roost, and <strong>for</strong>aging sites.<br />

5. Tree scale - use of individual roost and<br />

nest trees.<br />

In some cases, we reanalyzed existing data<br />

sets to provide more detailed in<strong>for</strong>mation or to<br />

explore different questions than those asked by<br />

original investigators. Methods varied among<br />

studies, and will be discussed in conjunction<br />

with <strong>the</strong> specific data examined. Some of <strong>the</strong><br />

data used were from ongoing studies, and some<br />

of <strong>the</strong> analyses are preliminary. There<strong>for</strong>e, while<br />

<strong>the</strong> in<strong>for</strong>mation presented here represents <strong>the</strong><br />

current state of knowledge on habitat relationships<br />

of <strong>Mexican</strong> spotted owls, we caution that<br />

additional trends may emerge with fur<strong>the</strong>r data<br />

collection and analysis.


PATTERNS PATTERNS OF OF HABITAT HABITAT USE USE AT<br />

AT<br />

THE THE LANDSCAPE LANDSCAPE SCALE<br />

SCALE<br />

Literature Literature Literature Review<br />

Review<br />

Available historical in<strong>for</strong>mation on habitat<br />

use by <strong>Mexican</strong> spotted owls in Arizona and<br />

New Mexico was reviewed in Johnson and<br />

Johnson (1985), Ganey (1988), Ganey et al.<br />

(1988), and Skaggs (1988), and summarized<br />

across <strong>the</strong> range of <strong>the</strong> owl in McDonald et al.<br />

(1991). Most reports of <strong>Mexican</strong> spotted owls in<br />

both popular and technical literature are anecdotal<br />

and contain little detailed in<strong>for</strong>mation.<br />

These reports were typically results of faunal<br />

surveys that reported areas where <strong>the</strong>y found<br />

species. As such, <strong>the</strong>y were nei<strong>the</strong>r systematic<br />

nor complete samples, and are thus biased<br />

towards <strong>the</strong> few places where spotted owls were<br />

located. While <strong>the</strong>se reports provide some<br />

in<strong>for</strong>mation on areas where owls were located,<br />

<strong>the</strong>y should not be viewed as <strong>the</strong> ultimate word<br />

on habitats used by spotted owls.<br />

With <strong>the</strong>se cautions in mind, <strong>the</strong> general<br />

picture that emerges is that owls occurred in<br />

habitats ranging from low elevation riparian<br />

<strong>for</strong>ests (Bendire 1892, Phillips et al. 1964, and<br />

possibly Woodhouse 1853:63) to high elevation<br />

coniferous <strong>for</strong>ests, but were most common in<br />

high elevation coniferous and mixed coniferousbroadleaved<br />

<strong>for</strong>ests, often in canyons. In perhaps<br />

<strong>the</strong> most detailed historical account of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl, Ligon (1926:422) described<br />

its typical haunts as “deep, narrow,<br />

timbered canyons where <strong>the</strong>re are always cool<br />

shady places.”<br />

Several recent studies describe habitats used<br />

by <strong>Mexican</strong> spotted owls. For example, Kertell<br />

(1977), Rinkevich (1991), and Willey (1992)<br />

reported on habitats occupied by <strong>Mexican</strong><br />

spotted owls in sou<strong>the</strong>rn Utah (Colorado Plateau<br />

RU). All reported that owls were typically found<br />

in narrow, steep-walled canyons, and all suggested<br />

that distribution was restricted by <strong>the</strong><br />

availability of such canyons. Reynolds (1993)<br />

also reported finding owls only in “steep-walled,<br />

deeply-cut canyons characterized or dominated<br />

by exposed rocky slopes and tiers of rock cliffs”<br />

in Colorado (Colorado Plateau and Sou<strong>the</strong>rn<br />

Volume II/Chapter 4<br />

2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Rocky Mountains-Colorado <strong>Recovery</strong> Units; see<br />

USDI 1995 <strong>for</strong> discussion of specific <strong>Recovery</strong><br />

Units [RUs] within <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl).<br />

Ganey and Balda (1989a) recorded cover<br />

type at 55 roost sites in nor<strong>the</strong>rn Arizona, of<br />

which 53 were located in <strong>the</strong> Upper Gila Mountains<br />

RU. Of <strong>the</strong>se, 92.4% were located in<br />

mixed-conifer <strong>for</strong>est. The remaining 7.6% were<br />

classified as ponderosa pine, but more likely were<br />

in ponderosa pine-Gambel oak <strong>for</strong>est. Most were<br />

located in steep canyons or on montane slopes.<br />

Also in this RU, Seamans and Gutiérrez (in<br />

press) recorded cover type at 79 roost and 28<br />

nest sites in <strong>the</strong> Tularosa Mountains, New<br />

Mexico. These owls roosted and nested primarily<br />

in mixed-conifer <strong>for</strong>ests containing an oak<br />

component, usually on <strong>the</strong> lower third of northfacing<br />

slopes (Seamans and Gutiérrez in press:<br />

fig. 1).<br />

In <strong>the</strong> Basin and Range-West RU, Ganey<br />

and Balda (1989a) recorded cover type at 64<br />

roost sites. Most were in mixed-conifer (48.4%)<br />

or Madrean pine-oak (29.7%) <strong>for</strong>est, with<br />

14.1% in encinal (evergreen oak), and 7.8% in<br />

ponderosa pine <strong>for</strong>est. At 19 roost and/or nest<br />

sites observed by Duncan and Taiz (1992) in this<br />

RU, 31.6% were in mixed-conifer <strong>for</strong>est, 31.6%<br />

in Madrean pine-oak <strong>for</strong>est, 26.3% in Arizona<br />

cypress <strong>for</strong>est, and 10.5% in encinal. In both<br />

studies, most owls were observed in montane<br />

canyons.<br />

Skaggs and Raitt (1988) surveyed 42,105 ha<br />

(104,000 acres) in <strong>the</strong> Basin and Range-East<br />

RU. Survey areas were divided into 18 plots of<br />

23.3-km 2 (9-mi 2 ) prior to survey, with all plots<br />

classified by <strong>the</strong> dominant <strong>for</strong>est type within <strong>the</strong><br />

plot (6 each in mixed-conifer, ponderosa pine, or<br />

pinyon-juniper). They found 33 occupied sites;<br />

72.7% in mixed-conifer, 18.2% in ponderosa<br />

pine, and 9.1% in pinyon-juniper. Even in plots<br />

classified as ponderosa pine or pinyon-juniper,<br />

<strong>the</strong> owls typically roosted in pockets of mixedconifer<br />

<strong>for</strong>est (Roger Skaggs, New Mexico State<br />

Univ., Las Cruces, NM, pers. comm.). Kroel<br />

(1991:39) also noted that owls in this RU<br />

roosted primarily (79% of observed roosts) in<br />

mixed-conifer <strong>for</strong>est, with limited use of ponderosa<br />

pine <strong>for</strong>est and pinyon-juniper woodland<br />

(14 and 4% of all roosts, respectively).


Little recent in<strong>for</strong>mation exists regarding<br />

habitat use by spotted owls in Mexico. However,<br />

Tarango et al. (1994) reported finding <strong>Mexican</strong><br />

spotted owls in isolated patches of pine-oak<br />

<strong>for</strong>est in steep canyons in <strong>the</strong> Sierra Madre<br />

Occidental-Norte.<br />

With regard to all of <strong>the</strong> above studies, it is<br />

important to note that definitions of cover types<br />

may have varied slightly among observers, and<br />

we cannot be certain that cover types are completely<br />

consistent with definitions used in this<br />

<strong>Plan</strong>.<br />

<strong>Recovery</strong> <strong>Recovery</strong> Team Team Analysis Analysis of of Recent<br />

Recent<br />

Inventory Inventory Data<br />

Data<br />

Recent inventories by land management<br />

agencies, particularly <strong>the</strong> U. S. Forest Service<br />

(FS), have generated considerable in<strong>for</strong>mation<br />

on owl locations and habitats used. We used this<br />

in<strong>for</strong>mation to evaluate use of vegetation (or<br />

cover) types across <strong>the</strong> range of <strong>the</strong> subspecies.<br />

Crews visited FS District and Supervisors offices,<br />

collated inventory and monitoring data from<br />

1990 (when survey ef<strong>for</strong>ts were standardized<br />

throughout <strong>the</strong> Region) to 1993 (when <strong>the</strong>se<br />

data were collated), and entered <strong>the</strong> in<strong>for</strong>mation<br />

ga<strong>the</strong>red into a data base. These same crews also<br />

collated records from <strong>the</strong> NPS, BLM, Tribal<br />

lands, State wildlife agencies, Fort Huachuca<br />

Military Reservation, and independent researchers<br />

in both <strong>the</strong> United States and Mexico <strong>for</strong> <strong>the</strong><br />

same time period. Thus, we attempted to ga<strong>the</strong>r<br />

all existing in<strong>for</strong>mation on current distribution<br />

and habitat use of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

across its range.<br />

Vegetation type at nest or roost sites was<br />

entered directly from field data <strong>for</strong>ms, with types<br />

generally corresponding to Series level designations<br />

described in Brown et al. (1980). We could<br />

not assess <strong>the</strong> accuracy or consistency of habitat<br />

classification across <strong>the</strong> range of <strong>the</strong> owl. All<br />

locations <strong>for</strong> which vegetation type in<strong>for</strong>mation<br />

was missing or ambiguous were omitted from<br />

analysis.<br />

We used only visual observations (such as<br />

birds at roost and nest sites) to assess habitat use<br />

because habitat cannot be determined <strong>for</strong> distant<br />

owls heard at night. Most roost or nest locations<br />

Volume II/Chapter 4<br />

3<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

entered in this database could not be assigned to<br />

a particular “management territory.” As a result,<br />

we are uncertain how many unique pairs of owls<br />

were represented, or how many roost or nest<br />

sites might represent a single pair. This potentially<br />

serious lack of independence in <strong>the</strong> data<br />

rendered statistical comparisons among RUs or<br />

between roost and nest sites meaningless. There<strong>for</strong>e,<br />

we simply present summaries of vegetation<br />

types used <strong>for</strong> roosting and nesting by <strong>Recovery</strong><br />

Unit. We could not compare habitat use with<br />

habitat occurrence at this scale, because of <strong>the</strong><br />

problems discussed above regarding lack of<br />

independence, because no geographical in<strong>for</strong>mation<br />

system (GIS) coverage was available documenting<br />

areas surveyed, and because we were<br />

unable to obtain a rangewide vegetation type<br />

coverage. The closest we could come to a<br />

rangewide vegetation type coverage covered only<br />

<strong>the</strong> U. S. portion of <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, and had a minimum resolution of 1<br />

km 2 (Keitt et al. 1995). This scale is not appropriate<br />

<strong>for</strong> evaluating roost and nest sites, which<br />

require a much finer scale of resolution.<br />

<strong>Mexican</strong> spotted owls used a variety of<br />

vegetation types across <strong>the</strong>ir range (Table 4.1).<br />

Although <strong>the</strong> range of vegetation types used<br />

varied among RUs, mixed-conifer <strong>for</strong>est was<br />

heavily used in most RUs. In contrast, encinal<br />

was used only in <strong>the</strong> Basin and Range-West, and<br />

pine-oak <strong>for</strong>est was used primarily in <strong>the</strong> Basin<br />

and Range-West and Sierra Madre Occidental-<br />

Norte. The pine-oak <strong>for</strong>est found in <strong>the</strong>se RUs is<br />

Madrean in affinity and differs in both species<br />

composition and habitat structure from <strong>the</strong><br />

ponderosa pine-Gambel oak <strong>for</strong>est found in<br />

o<strong>the</strong>r RUs (Brown et al. 1980, Ganey et al.<br />

1992). Ponderosa pine <strong>for</strong>est was used rarely, and<br />

pinyon-juniper woodland was used primarily by<br />

owls in <strong>the</strong> Colorado Plateau. Riparian <strong>for</strong>est<br />

was used in several RUs, but at relatively low<br />

levels.<br />

We could not statistically compare RUs<br />

because of <strong>the</strong> problems discussed above. Based<br />

on discussions with researchers and managers<br />

familiar with local situations, however, we<br />

believe that <strong>the</strong> observed differences in patterns<br />

of habitat use among RUs are both real and<br />

ecologically important.


4<br />

Table Table 4.1. 4.1. Percent of nest (top row) and roost (bottom row) sites in various vegetation types in different <strong>Recovery</strong> Units. Based on analysis of inventory<br />

and monitoring data collected since 1990. Vegetation types follow Series in Brown et al. (1980).<br />

FPO


PATTERNS PATTERNS OF OF HABITAT HABITAT USE USE AT<br />

AT<br />

THE THE THE HOME-RANGE HOME-RANGE SCALE<br />

SCALE<br />

Home-Range Home-Range Size<br />

Size<br />

Several studies have examined home-range<br />

size and/or habitat use within home ranges of<br />

radio-tagged <strong>Mexican</strong> spotted owls. Minimum<br />

convex polygon (MCP; Mohr 1947) and 95%<br />

adaptive kernel (AK; Worton 1989, where<br />

possible) estimates of home-range size from <strong>the</strong>se<br />

studies are presented in Table 4.2. Estimates are<br />

presented separately <strong>for</strong> each study area. Homerange<br />

size appeared to vary considerably among<br />

study areas, even within a restricted geographic<br />

area (Table 4.2). For example, Zwank et al.<br />

(1994) found that home-range size differed<br />

significantly between two drainages in <strong>the</strong><br />

Sacramento Mountains, New Mexico. Ranges<br />

were smaller in <strong>the</strong> Rio Penasco watershed (n = 5<br />

owls), where mixed-conifer <strong>for</strong>est comprised 65-<br />

86% of home ranges, than in <strong>the</strong> Sixteen Springs<br />

watershed (n = 4 owls), where mixed-conifer<br />

<strong>for</strong>est comprised only 6-42% of home ranges,<br />

with most of <strong>the</strong> remainder consisting of ponderosa<br />

pine <strong>for</strong>est and pinyon-juniper woodland<br />

(Zwank et al. 1994: table 1). Similarly, Ganey<br />

and Balda (1989b) observed large differences in<br />

home-range size among owls on <strong>the</strong> San Francisco<br />

Peaks and in Walnut Canyon. Although<br />

<strong>the</strong>se areas are separated by only approximately<br />

16 km (10 mi), habitat composition differs<br />

between <strong>the</strong> two areas (Ganey and Balda 1994:<br />

table 3). These results suggest that home-range<br />

size of <strong>Mexican</strong> spotted owls may vary among<br />

cover types. However, small numbers of owls<br />

tracked in all studies, as well as differences in<br />

sampling intervals and seasons covered limit our<br />

ability to make comparisons among study areas.<br />

Consequently, we caution that <strong>the</strong>se numbers<br />

represent general estimates of areas used by<br />

spotted owls, and as such do not support sweeping<br />

generalizations about differences between<br />

areas and/or cover types.<br />

Size Size of of Activity Activity Centers<br />

Centers<br />

Gutiérrez et al. (1992) suggested that <strong>the</strong><br />

smallest area encompassing 50% of nocturnal<br />

Volume II/Chapter 4<br />

5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

<strong>for</strong>aging locations (<strong>the</strong> 50% adaptive kernel)<br />

could define a <strong>for</strong>aging activity center. Because<br />

of <strong>the</strong> great variability in available estimates of<br />

home-range size <strong>for</strong> <strong>Mexican</strong> spotted owls, we<br />

took a more conservative approach and defined a<br />

nocturnal activity center based on <strong>the</strong> area<br />

enclosed by <strong>the</strong> adaptive kernel contour encompassing<br />

75% of <strong>for</strong>aging locations. This activity<br />

center was defined only <strong>for</strong> territories where<br />

both pair members were radio-tagged. In general,<br />

owls appeared to <strong>for</strong>age primarily within a<br />

relatively small portion of <strong>the</strong> home range,<br />

suggesting high concentration of activity (Table<br />

4.3). This pattern appeared to hold regardless of<br />

RU or study area (but see above cautions regarding<br />

comparisons among studies).<br />

Habitat Habitat Composition Composition and and Use<br />

Use<br />

Three studies quantified habitat composition<br />

within MCP home ranges of radio-tagged owls<br />

(Willey 1993, Ganey and Balda 1994, Zwank et<br />

al. 1994). Home ranges were most variable in<br />

terms of vegetation types in <strong>the</strong> Colorado<br />

Plateau RU, encompassing types ranging from<br />

mixed-conifer <strong>for</strong>est to mountain shrub and<br />

grassland (Willey 1993: table 4). Home ranges<br />

were dominated by mixed-conifer and ponderosa<br />

pine <strong>for</strong>ests in <strong>the</strong> Upper Gila Mountains RU<br />

(Ganey and Balda 1994), and by mixed-conifer<br />

<strong>for</strong>est, ponderosa pine <strong>for</strong>est, and pinyon-juniper<br />

woodland in <strong>the</strong> Basin and Range-East RU<br />

(Zwank et al. 1994).<br />

Ganey and Balda (1994) compared use of<br />

vegetation types by <strong>for</strong>aging, radio-tagged owls<br />

to <strong>the</strong> area of those types within <strong>the</strong> MCP home<br />

range (a measure of relative availability), using<br />

chi-square tests and Bonferroni confidence<br />

intervals (Neu et al. 1974, Byers et al. 1984).<br />

This comparison involved eight owls representing<br />

five pairs on three study areas.<br />

Observed patterns of habitat use by <strong>for</strong>aging<br />

owls were complex. In relation to area of different<br />

<strong>for</strong>est types within <strong>the</strong>ir home ranges, all<br />

individual owls used <strong>for</strong>est types nonrandomly.<br />

All <strong>for</strong>est types were used by <strong>for</strong>aging owls. In<br />

general, individual owls <strong>for</strong>aged significantly<br />

more than or as expected in unlogged <strong>for</strong>ests and<br />

significantly less than or as expected in selec-


6<br />

Table Table 4.2. 4.2. Home-range sizes (ha) of radio-marked <strong>Mexican</strong> spotted owls. Part A shows ranges of pairs with both members radio-tagged; part B shows<br />

ranges <strong>for</strong> individual owls. Study areas represent mesic mixed-conifer <strong>for</strong>est on <strong>the</strong> San Francisco Peaks (SFP), White Mountains (WM), and Sacramento<br />

Mountains (SM-MC); a mixture of mixed-conifer, ponderosa pine, and xeric pinyon-juniper in <strong>the</strong> Sacramento Mountains (SM-XE); a rocky canyon<br />

(Walnut Canyon, WC); ponderosa pine-Gambel oak <strong>for</strong>est near Bar-M Canyon (BMC); rugged canyons with mixed <strong>for</strong>ests in <strong>the</strong> San Mateo Mountains<br />

(SANMAT); a mixture of mesic mixed-conifer <strong>for</strong>est and pinyon-juniper in rocky canyons along Elk Ridge (MANTI) and in Zion National Park (ZION);<br />

and xeric pinyon-juniper in rocky canyons in Capitol Reef and Canyonlands National Parks (CNP). Shown are <strong>the</strong> means (± SD) <strong>for</strong> 100% minimum<br />

convex polygon (MCP) and 95% adaptive kernel (95% AK) estimates.<br />

FPO


7<br />

Table Table 4.2. 4.2. (continued)<br />

FPO


8<br />

Table Table 4.3. 4.3. Size (mean ± standard deviation) of nocturnal activity centers of radio-tagged pairs of Mexcan spotted owls in <strong>the</strong> Upper Gila Mountains and<br />

Basin and Range-East <strong>Recovery</strong> units. Activity centers defined as <strong>the</strong> area included in <strong>the</strong> adaptive kernel contour enclosing 75% of owl <strong>for</strong>aging locations.<br />

FPO


tively logged <strong>for</strong>est types (Table 4.4). This<br />

comparison is overly simplistic, however. Both<br />

logged and unlogged <strong>for</strong>ests contained a wide<br />

range of stand structures. Some logged stands<br />

were used <strong>for</strong> <strong>for</strong>aging, and some unlogged<br />

stands were not used. In addition, two of <strong>the</strong><br />

unlogged <strong>for</strong>est types (virgin mixed-conifer<br />

<strong>for</strong>est on rocky slopes and ponderosa pine-oakjuniper<br />

<strong>for</strong>est) were found primarily on rocky<br />

slopes interspersed with significant rock outcrops<br />

and cliffs, and owls appeared to <strong>for</strong>age in <strong>the</strong>se<br />

rocky areas as well as in <strong>the</strong> <strong>for</strong>est (Ganey and<br />

Balda (1994:165). Thus, <strong>the</strong>se findings do not<br />

indicate that all unlogged stands provide suitable<br />

<strong>for</strong>aging conditions and that all logged stands do<br />

not. Ra<strong>the</strong>r, <strong>the</strong>y suggest that patterns of habitat<br />

use by <strong>for</strong>aging owls are complex and not amenable<br />

to simplistic explanations. At this time, we<br />

cannot explain why some logged <strong>for</strong>ests are used<br />

and o<strong>the</strong>rs are not.<br />

Differences in patterns of habitat use among<br />

and within study areas also suggest that patterns<br />

of habitat use <strong>for</strong> <strong>for</strong>aging are complex. Use of<br />

particular <strong>for</strong>est types sometimes varied considerably<br />

among individual owls within a study area<br />

(Table 4.4), and even between members of a<br />

mated pair (Ganey and Balda 1994: table 3). In<br />

short, <strong>the</strong>re is considerable variability in patterns<br />

of habitat use <strong>for</strong> <strong>for</strong>aging. As a result, it is<br />

difficult to generalize about patterns of habitat<br />

selection by <strong>for</strong>aging owls based on currentlyavailable<br />

data.<br />

In contrast to <strong>the</strong> variable patterns observed<br />

among <strong>for</strong>aging owls, patterns of habitat use by<br />

roosting owls were relatively consistent among<br />

study areas and individual owls. Habitat use <strong>for</strong><br />

roosting was not compared statistically to habitat<br />

occurrence because of small sample sizes <strong>for</strong><br />

some individual owls. All owls roosted primarily<br />

in unlogged mixed-conifer <strong>for</strong>ests, but some<br />

used unlogged ponderosa pine <strong>for</strong>est as well<br />

(Ganey and Balda 1994; table 3). Very little<br />

roosting occurred in logged stands, particularly<br />

during <strong>the</strong> breeding season.<br />

Volume II/Chapter 4<br />

9<br />

Seasonal Seasonal Movements Movements<br />

Movements<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Most <strong>Mexican</strong> spotted owls appear to remain<br />

in <strong>the</strong> same general area throughout <strong>the</strong> year,<br />

whereas o<strong>the</strong>rs migrate in winter, usually to<br />

lower elevations. Year-round residents often use<br />

larger ranges during <strong>the</strong> nonbreeding season<br />

than during <strong>the</strong> breeding season (Kroel 1991,<br />

Willey 1993, Ganey and Block unpublished<br />

data), and <strong>the</strong>re appear to be shifts in use of area<br />

and habitat <strong>for</strong> some owls (Ganey and Balda<br />

1989b, Kroel 1991, Willey 1993, Ganey and<br />

Block unpublished data). No quantitative results<br />

are yet available that describe wintering habitats<br />

used by owls remaining in <strong>the</strong> same area<br />

throughout <strong>the</strong> year.<br />

Most samples of <strong>Mexican</strong> spotted owls<br />

studied using radiotelemetry appear to have<br />

some individuals who are ei<strong>the</strong>r migratory or<br />

nomadic in winter, but this generally represents a<br />

minority of <strong>the</strong> population. For example, Willey<br />

(1993:15) reported that two of 11 (18.2%)<br />

radio-tagged owls on <strong>the</strong> Colorado Plateau RU<br />

migrated during winter. Both apparently left <strong>the</strong><br />

breeding area during October and returned<br />

during February. One moved up in elevation to<br />

winter in coniferous <strong>for</strong>est, whereas <strong>the</strong> o<strong>the</strong>r<br />

wintered in mountain shrub habitat. Both owls<br />

moved 20-25 km between breeding season<br />

ranges and winter ranges.<br />

In <strong>the</strong> Upper Gila Mountains RU, two of<br />

eight owls (25%) in one study (Ganey and Balda<br />

1989b) and 2 of 13 (15.4%) in ano<strong>the</strong>r (Ganey<br />

and Block unpublished data) migrated during<br />

<strong>the</strong> winter. All left <strong>the</strong> breeding-season range<br />

between November and January, and returned in<br />

March or April. Wintering areas of two owls<br />

were never located. The remaining two owls<br />

migrated approximately 50 km, from ponderosa<br />

pine-Gambel oak <strong>for</strong>est at approximately 2290<br />

m (7500 ft) in elevation to pinyon-juniper<br />

woodland at approximately 1370 m (4490 ft) in<br />

elevation (Ganey et al. 1992, Ganey and Block<br />

unpublished data). The wintering area was<br />

located in <strong>the</strong> Basin and Range-West RU,<br />

providing evidence that some individuals may<br />

move seasonally between RUs.


10<br />

Table Table 4.4. 4.4. Use of <strong>for</strong>est types <strong>for</strong> <strong>for</strong>aging by radio-tagged <strong>Mexican</strong> spotted owls on three study areas in nor<strong>the</strong>rn Arizona. Data summarized from<br />

Ganey and Balda (1994).<br />

FPO


In <strong>the</strong> Basin and Range-East RU, nei<strong>the</strong>r<br />

Zwank et al. (1994, n = 9 owls) nor Ganey and<br />

Block (unpublished data, n = 15 owls) observed<br />

migration during <strong>the</strong> nonbreeding season, but<br />

two of eight radio-tagged owls (25%) in ano<strong>the</strong>r<br />

study moved during <strong>the</strong> winter. These owls<br />

moved downslope from mixed-conifer <strong>for</strong>ests to<br />

<strong>the</strong> interface between pinyon-juniper woodland<br />

and desert scrub (Roger Skaggs, New Mexico<br />

State Univ., Las Cruces, NM; pers. comm.).<br />

PATTERNS PATTERNS OF OF HABITAT HABITAT USE USE AT<br />

AT<br />

THE THE STAND STAND SCALE<br />

SCALE<br />

Limited data were available on habitat use by<br />

<strong>Mexican</strong> spotted owls at <strong>the</strong> stand scale during<br />

preparation of this <strong>Recovery</strong> <strong>Plan</strong>. Analysis of FS<br />

data <strong>for</strong> some nesting stands in <strong>the</strong> Upper Gila<br />

Mountains and Basin and Range-East RUs<br />

suggests that owls typically nest in relatively<br />

dense stands with high basal areas of live trees, a<br />

FPO<br />

wide range of tree sizes, suggesting an unevenaged<br />

structure (Figure 4.1), and a large tree<br />

component (Table 4.5). The data were also used<br />

to estimate diminution quotients, or q-factors,<br />

<strong>for</strong> nesting stands in both RUs (Table 4.5). This<br />

parameter is useful in uneven-aged management<br />

of timber stands. It describes <strong>the</strong> ratio of number<br />

of trees in any diameter class to <strong>the</strong> number in<br />

<strong>the</strong> next-lowest diameter class, and thus describes<br />

<strong>the</strong> relative shape of <strong>the</strong> diameter distribution<br />

(Daniel et al. 1979). In both RUs, q-factors<br />

averaged


12<br />

Table Table 4.5. 4.5. Selected characteristics of nest stands of <strong>Mexican</strong> spotted owls in <strong>the</strong> Upper Gila Mountains (n = 13 stands) and Basin and Range-East<br />

(n = 44 stands) <strong>Recovery</strong> Units. Data from <strong>the</strong> stand data bases <strong>for</strong> <strong>the</strong> Apache-Sitgreaves (Upper Gila Mountains) and Lincoln (Basin and Range-East)<br />

National Forests. Values shown are means; no estimates of variability among stands were available.<br />

FPO


Nest Nest Sites Sites<br />

Sites<br />

Armstrong et al. (1994) and Arizona Game<br />

and Fish Department (unpublished data)<br />

sampled habitat characteristics at nest sites on<br />

<strong>the</strong> Apache-Sitgreaves National Forest, Upper<br />

Gila Mountains RU (Table 4.6). They sampled<br />

both tree and cliff nests, but <strong>the</strong> majority of<br />

nests sampled were in trees. While limited in<br />

scope, <strong>the</strong>ir results suggest that owls typically<br />

nested in large trees in closed-canopy stands.<br />

Ruess (1995) documented current stand<br />

structure on 0.04-ha (0.1-ac) plots at 11 nest<br />

and 9 roost sites representing an unknown<br />

number of owl pairs in ponderosa pine-Gambel<br />

oak <strong>for</strong>est in northcentral Arizona. He also<br />

attempted to estimate what <strong>the</strong>se sites would<br />

have looked like in terms of <strong>for</strong>est structure in<br />

1876, be<strong>for</strong>e effective fire suppression began in<br />

this area, using methods developed by<br />

Covington and Moore (1994). With a few<br />

exceptions, Ruess (1995) pooled roost and nest<br />

sites when presenting data. Thus, both site types<br />

will be discussed toge<strong>the</strong>r here.<br />

Comparisons of current and estimated<br />

presettlement conditions on <strong>the</strong>se sites suggest<br />

that pronounced increases have occurred in tree<br />

density, basal area, and canopy cover. Current<br />

density of trees above breast height averaged<br />

1708.8 ± 1009.1 (SD) trees/ha (691.8 ± 408.5<br />

trees/ac), versus an estimated presettlement<br />

density of 0-225 trees/ha (0-91 trees/ac; [Ruess<br />

1995:12, no mean presented]). Current basal<br />

area averaged 66.7 m 2 /ha (290.7 ft 2 /ac; [Ruess<br />

1995:12, no estimate of variability provided]),<br />

versus an estimated average basal area of 17.8<br />

m 2 /ha (77.6 ft 2 /ac; [Ruess 1995:12, no estimate<br />

of variability provided]) circa 1876. Canopy<br />

cover, modeled based on projection of mapped<br />

tree crowns, was estimated at 44.8 ± 12.9% at<br />

present, versus 2.2 ± 2.9% circa 1876 (Ruess<br />

1995:22). Variability in species composition and<br />

structural variables was relatively high <strong>for</strong> both<br />

estimated 1876 conditions and current conditions.<br />

Two detailed treatments of nesting habitat of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl are currently available<br />

(SWCA 1992, Seamans and Gutiérrez in press).<br />

Seamans and Gutiérrez (in press) compared<br />

Volume II/Chapter 4<br />

13<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

habitat characteristics between 27 plots (0.04 ha;<br />

0.1 ac) centered on nest trees and 27 random<br />

plots from throughout <strong>the</strong>ir study area (Tularosa<br />

Mountains, Gila National Forest, New Mexico;<br />

Upper Gila Mountains RU). The nest plots<br />

represented nest sites of 27 pairs of owls. Random<br />

plots were centered on a randomly-selected<br />

tree ³27.3 cm diameter at breast height (dbh),<br />

<strong>the</strong> minimum diameter among <strong>the</strong> 27 nest trees.<br />

This was an attempt to minimize <strong>the</strong> potential<br />

bias associated with centering nest plots on large<br />

trees and random plots on trees of any size.<br />

<strong>Owl</strong>s nested in mixed-conifer/oak <strong>for</strong>ests<br />

more than expected by chance, and in pine-oak<br />

and pinyon-juniper <strong>for</strong>ests less than expected<br />

(Seamans and Gutiérrez in press: fig. 1). Most<br />

nests were located on <strong>the</strong> lower third of slopes,<br />

and <strong>the</strong> mean slope aspect at nest sites was<br />

nor<strong>the</strong>rly. Nest plots differed significantly from<br />

randomly-located plots within <strong>the</strong> study area <strong>for</strong><br />

a number of variables (Table 4.7). In a discriminant<br />

function analysis, nest plots were best<br />

separated from random plots by variance in tree<br />

height, canopy closure, and basal area of mature<br />

trees (defined as stems >45.8 cm dbh); all were<br />

greater on nest than on random plots (Table<br />

4.7). Cross-validation analyses indicated that <strong>the</strong><br />

results of <strong>the</strong> discriminant analysis were stable,<br />

and <strong>the</strong> discriminant function successfully<br />

classified 84.6% of a sample of 13 owl nest sites<br />

from o<strong>the</strong>r mountain ranges outside <strong>the</strong> study<br />

area (Seamans and Gutiérrez in press).<br />

Seamans and Gutiérrez (in press) also compared<br />

nest plots to 27 plots randomly located<br />

within nest stands. Nest plots did not differ<br />

significantly from random plots within <strong>the</strong> same<br />

stand.<br />

SWCA (1992) sampled habitat characteristics<br />

at 84 nests on FS lands in Arizona and New<br />

Mexico. They sampled habitat characteristics<br />

within circular plots (0.2 ha; 0.5 ac), each<br />

centered on and including ei<strong>the</strong>r a nest tree, a<br />

randomly selected tree within <strong>the</strong> nest stand, or<br />

a randomly selected tree in a stand within 0.8<br />

km (0.5 mi) of <strong>the</strong> nest. These will be referred to<br />

as nest, nest-stand, and random-stand plots.<br />

SWCA (1992) concluded that owls selected nest<br />

sites based primarily on <strong>the</strong> availability of a<br />

suitable nest tree. Hardwood snag basal area and<br />

canopy cover also emerged as potentially impor-


14<br />

Table Table 4.6. 4.6. Habitat characteristics sampled at <strong>Mexican</strong> spotted owl nest sites on three Ranger districts, Apache-Sitgreaves National Forest, Upper Gila<br />

Mountains <strong>Recovery</strong> Unit. Shown are means and standard errors (in paren<strong>the</strong>ses). Nests sampled were found in trees (n = 30) or on cliffs (n = 4); some<br />

variables are relevant to only one of <strong>the</strong> two situations.<br />

FPO


Table Table 4.7. 4.7. Habitat characteristics at <strong>Mexican</strong> spotted owl nest (n = 27) and randomly located sites<br />

(n = 27) in <strong>the</strong> Tularosa Mountains, New Mexico; Upper Gila Mountains RU. Shown are means and<br />

standard deviation (in paren<strong>the</strong>ses). Data from Seamans and Gutiérrez (in press).<br />

tant factors in <strong>the</strong>ir analysis. This data set was<br />

reanalyzed by both Zhou (1994) and <strong>the</strong> Team,<br />

using different methods. These reanalyses are<br />

discussed below.<br />

Reanalysis Reanalysis Reanalysis of of SWCA SWCA (1992) (1992) by by by Zhou Zhou Zhou (1994)<br />

(1994)<br />

Zhou (1994) conducted three, two-group<br />

linear discriminant function analyses between<br />

nest, nest-stand, and random-stand plots. He<br />

concluded (Zhou 1994:96-102) that:<br />

1. <strong>Mexican</strong> spotted owl nest stands differed<br />

from randomly-selected stands in <strong>the</strong><br />

vicinity. Nest stands typically had a wide<br />

range of tree diameters and heights, large<br />

maximum tree diameter, and high tree<br />

basal area. Nest stands also had higher<br />

species richness than random stands.<br />

2. <strong>Mexican</strong> spotted owls also selected <strong>for</strong><br />

microsites within nest stands. Nest plots<br />

were located on steeper slopes, had<br />

greater live tree basal area, and were more<br />

likely to be found on north or east<br />

aspects than nest-stand plots.<br />

Volume II/Chapter 4<br />

FPO<br />

15<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

3. Diameter and height distributions had<br />

similar shapes on all three plot types, but<br />

<strong>the</strong> spread of <strong>the</strong> distribution differed<br />

among plot types. With respect to<br />

diameter distributions, <strong>the</strong> nest and neststand<br />

plots had greater percentages of<br />

large trees. With respect to tree height,<br />

nest and nest-stand plots had greater<br />

spread to <strong>the</strong> distribution than randomstand<br />

plots. This wider spread suggests a<br />

tendency <strong>for</strong> <strong>the</strong> owl to nest in multistoried<br />

stands, as wider spread to <strong>the</strong><br />

height distribution increases <strong>the</strong> probability<br />

that <strong>the</strong> stand is multi-storied.<br />

4. <strong>Mexican</strong> spotted owls nested in large<br />

trees. The mean nest tree was located in<br />

<strong>the</strong> 92nd diameter percentile and <strong>the</strong><br />

79th height percentile.<br />

5. The linear combination of habitat<br />

variables was more successful in classify<br />

ing habitat than reliance on interpretation<br />

of coefficients <strong>for</strong> single variables.


Reanalysis Reanalysis of of SWCA SWCA (1992) (1992) by by <strong>the</strong><br />

<strong>the</strong><br />

<strong>Recovery</strong> <strong>Recovery</strong> Team<br />

Team<br />

The Team also reanalyzed <strong>the</strong> data from<br />

SWCA (1992). The Team was interested in<br />

patterns of habitat use within particular geographic<br />

regions, and in how similar nest sites<br />

were among regions. This reanalysis was restricted<br />

to sites in <strong>the</strong> Upper Gila Mountains<br />

and Basin and Range-East RUs because <strong>the</strong>se<br />

were <strong>the</strong> only RUs well represented among nests<br />

sampled (n = 44 and 26 sites, respectively).<br />

Because sample sizes were small, sites were<br />

pooled among habitat types within each RU.<br />

Habitat characteristics were first compared<br />

among plot types within RUs, to see if nest sites<br />

differed from nest stands or locally-available,<br />

randomly-selected stands. Characteristics of nest<br />

plots were next compared between RUs, to see<br />

how similar nesting habitat was in different<br />

geographic areas.<br />

We used chi-square tests to evaluate differences<br />

in tree species composition between plot<br />

types (and/or RUs), and Kolmogorov-Smirnov<br />

tests to compare diameter distributions. Because<br />

<strong>the</strong> Kolmogorov-Smirnov test compares only<br />

two groups at a time, three separate tests were<br />

necessary to compare all three plot types. To<br />

avoid inflating <strong>the</strong> Type I error rate, we partitioned<br />

<strong>the</strong> error among <strong>the</strong>se three comparisons<br />

using a Bonferroni adjustment, and used an<br />

alpha level of P 12 cm (4.7 in) dbh, basal area<br />

(m 2 /ha) of live trees and snags >12 cm dbh, and<br />

volume of logs (m 3 /ha) >12 cm in large-enddiameter.<br />

Basal area was calculated <strong>for</strong> each<br />

individual tree or snag based on dbh, <strong>the</strong>n<br />

summed within each plot. Log volume was<br />

calculated using diameter and length measures,<br />

assuming a cylindrical shape. Log volumes are<br />

overestimated (perhaps greatly) because diameter<br />

was measured at <strong>the</strong> large end.<br />

Volume II/Chapter 4<br />

16<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Comparisons omparisons Within ithin ithin R RReco<br />

R Reco<br />

eco ecover eco er ery ery<br />

y U UUnits.—<br />

U nits.— nits.—Tree nits.—<br />

species composition differed significantly among<br />

plot types in both RUs. In <strong>the</strong> Upper Gila<br />

Mountains RU, nest plots contained greater<br />

proportions of Douglas-fir and Gambel oak and<br />

less ponderosa pine than did random-stand plots<br />

(Figure 4.2). In <strong>the</strong> Basin and Range-East RU,<br />

nest and nest-stand plots contained more white<br />

fir and less ponderosa pine than did randomstand<br />

plots (Figure 4.2).<br />

Diameter distributions in both RUs were<br />

significantly different between nest plots and<br />

both nest-stand and random-stand plots, but<br />

were not significantly different between neststand<br />

and random-stand plots. Nest plots typically<br />

had lower proportions of basal area in <strong>the</strong><br />

smallest size classes than did random-stand plots<br />

(Figure 4.3). In general, basal area was more<br />

evenly distributed across size classes in nest<br />

stands than in random stands, suggesting a trend<br />

toward uneven-aged stands. This trend was more<br />

evident in <strong>the</strong> Basin and Range-East RU than in<br />

<strong>the</strong> Upper Gila Mountains RU. Grouping of<br />

trees into four diameter classes, representing<br />

“young,” “mid-aged,” “mature,” and “old” trees<br />

(USDA Forest Service 1993), also indicates that<br />

nest plots contained relatively fewer trees in <strong>the</strong><br />

smallest size class and more trees in <strong>the</strong> largest<br />

two size classes than o<strong>the</strong>r plots (Table 4.8a, b;<br />

note that <strong>the</strong>se comparisons refer to percentages<br />

of total trees, not to absolute numbers). In both<br />

RUs, nest trees were significantly larger<br />

(P


Volume II/Chapter 4<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 4.2a. 4.2a. Tree species composition within nest- and random-stand plots. Data reanalyzed from<br />

SWCA (1992). Upper Gila Mountains <strong>Recovery</strong> Unit (n = 44 sites).<br />

17


Figure Figure 4.2b. 4.2b. 4.2b. Basin and Range-East <strong>Recovery</strong> Unit (n = 26 sites).<br />

Volume II/Chapter 4<br />

FPO<br />

18<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Volume II/Chapter 4<br />

FPO<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 4.3. 4.3. Diameter distributions of live trees sampled on nest, nest stand, and random stand plots.<br />

Shown is percentage of total live tree basal area by 4 in (10 cm) size classes <strong>for</strong> (a) (a) Upper Gila Mountains,<br />

and (b) (b) Basin and Range-East. Data reanalyzed from SWCA (1992).<br />

19


log volume, both nest plots and nest-stand plots<br />

differed from random-stand plots, but not from<br />

each o<strong>the</strong>r (Table 4.8a).<br />

Of <strong>the</strong> five characteristics discussed above,<br />

only live tree basal area and log volume differed<br />

significantly (P = 0.0121 and 0.0061, respectively)<br />

between plot types in <strong>the</strong> Basin and<br />

Range-East RU (all o<strong>the</strong>r P values >0.05). Both<br />

variables differed only between nest plots and<br />

random-stand plots, not between nest plots and<br />

nest-stand plots or between nest-stand and<br />

random-stand plots (Table 4.8b).<br />

Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.8a. 4.8a. Habitat characteristics sampled at 44 <strong>Mexican</strong> spotted owl nest sites in <strong>the</strong> Upper Gila<br />

Mountains <strong>Recovery</strong> Unit, as well as at randomly located plots within <strong>the</strong> nest stand and in a randomly<br />

selected stand within 0.5 mi of <strong>the</strong> nest site. Data reanalyzed from SWCA (1992). Values shown are<br />

mean (± standard deviation).<br />

FPO<br />

20<br />

Comparisons omparisons Betw Between Betw een R RReco<br />

R eco ecover eco er ery er y UU<br />

Units.— UU<br />

nits.— nits.— nits.—Species<br />

nits.—<br />

composition of nest sites differed significantly<br />

between RUs (C2 = 1669, df = 4, P


Roost Roost sites sites<br />

sites<br />

Many studies have described characteristics<br />

of roost sites. Some of <strong>the</strong>se descriptions are<br />

based on plot-level sampling, whereas o<strong>the</strong>rs are<br />

based on sampling of microsites (such as an<br />

individual tree). Microsite descriptions will be<br />

discussed under “Patterns of Habitat Use at <strong>the</strong><br />

Tree scale”; plot-level data are discussed below.<br />

In most of <strong>the</strong> studies described here, <strong>the</strong><br />

number of plots measured exceeded <strong>the</strong> number<br />

of owls studied. Thus, <strong>the</strong>se plots cannot be<br />

considered totally independent samples, and<br />

Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.8b. 4.8b. 4.8b. Habitat characteristics sampled at 26 <strong>Mexican</strong> spotted owl nest sites in <strong>the</strong> Basin and<br />

Range-East <strong>Recovery</strong> Unit, as well as at randomly located plots within <strong>the</strong> nest stand and in a randomly<br />

selected stand within 0.5 mi of <strong>the</strong> nest site. Data reanalyzed from SWCA (1992). Values<br />

shown are mean (± standard deviation).<br />

FPO<br />

21<br />

apparent levels of significance may be inflated<br />

(Hurlbert 1984). In most cases, significance<br />

levels are so high that we believe pseudoreplication<br />

is not a major problem. This is fur<strong>the</strong>r<br />

suggested by <strong>the</strong> fact that results are comparable<br />

between <strong>the</strong>se studies and ano<strong>the</strong>r (Seamans and<br />

Gutiérrez in press) that did not involve pseudoreplication<br />

(see below).<br />

Rinkevich (1991; see also Rinkevich and<br />

Gutiérrez in review) and Willey (1993) sampled<br />

habitat characteristics within roost-centered<br />

circular plots of 0.04 ha (0.1 ac) each within <strong>the</strong><br />

Colorado Plateau RU. All roost sites were


located in narrow gorges and/or canyons.<br />

Rinkevich (1991) used discriminant function<br />

analysis to compare owl roost sites to randomly<br />

located sites in Zion National Park, Utah. <strong>Owl</strong><br />

sites had higher absolute humidity, more vegetation<br />

strata, narrower canyon width, and higher<br />

percent ground litter than random sites<br />

(Rinkevich and Gutiérrez in review). She also<br />

compared randomly located plots (n = 54)<br />

within canyons where owls were heard to plots<br />

(n = 44) within canyons where owls were not<br />

heard. Canyons occupied by owls had higher<br />

humidity and snag basal area than canyons<br />

where owls were not heard (Rinkevich and<br />

Gutiérrez in review). The number of plots in<br />

<strong>the</strong>se analyses is greater than <strong>the</strong> number of owls<br />

(or canyons in <strong>the</strong> second analysis), which is<br />

unknown.<br />

Willey (1993) sampled habitat characteristics<br />

on 129 plots representing roost sites of 14 radiotagged<br />

owls on three study areas in sou<strong>the</strong>rn<br />

Utah. Habitat features at roost sites were compared<br />

to characteristics measured at 30-50<br />

random points scattered within each home range<br />

(Willey 1993:10). Seven variables were analyzed<br />

using discriminant function analysis (Willey<br />

1993:10). Temperature, slope, vegetation canopy<br />

cover, and number of ledges and fir trees discriminated<br />

between roosting plots and random<br />

plots (Willey 1993: table 5). Univariate analyses<br />

provided similar results, suggesting that “owls<br />

used narrow canyon roosts characterized by cool<br />

daytime temperatures, steep slopes, and relatively<br />

dense overhead cover. Roosts typically possessed<br />

large trees in juxtaposition with caves and ledges,<br />

providing a more complex habitat architecture<br />

than <strong>the</strong> surrounding habitat” (Willey 1993:16).<br />

The number of plots also exceeded <strong>the</strong> number<br />

of owls in this study.<br />

Ganey (1988, see also Ganey and Balda<br />

1994) sampled habitat characteristics on 167<br />

circular plots (0.04 ha; 0.1 ac) within four owl<br />

home ranges (defined using <strong>the</strong> MCP method)<br />

in <strong>the</strong> Upper Gila Mountains RU. Plots represented<br />

high-use roosting and <strong>for</strong>aging sites, and<br />

randomly-selected sites within owl ranges. This<br />

study also had more plots than owls. In addition,<br />

roost plots were always tree-centered, whereas<br />

only some <strong>for</strong>aging and randomly-selected plots<br />

were tree-centered. This could create a positive<br />

Volume II/Chapter 4<br />

22<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

bias <strong>for</strong> tree-related variables, such as tree density,<br />

tree basal area, and canopy cover, on roost<br />

plots.<br />

Plot type was misclassified 33% of <strong>the</strong> time<br />

in a 3-group discriminant function analysis<br />

(Ganey 1988; all classification rates refer to<br />

jackknifed classification). Most misclassification<br />

occurred between <strong>for</strong>aging and roosting sites.<br />

Two-group analyses had higher rates of successful<br />

classification and were easier to interpret. The<br />

function that resulted in maximum separation of<br />

roosting and <strong>for</strong>aging sites correctly classified<br />

76% of <strong>the</strong> sites. Variables entering <strong>the</strong> equation<br />

were canopy closure and snags/ha; both were<br />

greater on roosting sites (Table 4.9). A comparison<br />

of <strong>for</strong>aging and random sites resulted in<br />

84% successful classification. Variables entering<br />

<strong>the</strong> discriminant function were total basal area<br />

and big down logs/ha (defined as logs >30.5 cm<br />

[12.0 in] in diameter); both were greater on<br />

<strong>for</strong>aging sites. Comparing roosting and random<br />

sites, 90% of all sites were successfully classified.<br />

Variables entering <strong>the</strong> discriminant function<br />

were total basal area, snags/ha, canopy closure,<br />

and big down logs/ha; all were greater on roosting<br />

sites.<br />

Using data from <strong>the</strong> plots sampled by Ganey<br />

(1988), <strong>the</strong> Team conducted Kolmogorov-<br />

Smirnov tests on diameter distributions as<br />

described under nest sites (see above). We found<br />

no difference in diameter distributions between<br />

roosting sites and ei<strong>the</strong>r <strong>for</strong>aging or random<br />

sites. Diameter distributions were significantly<br />

different between <strong>for</strong>aging and random sites. In<br />

general, <strong>for</strong>aging sites had fewer small trees and<br />

more trees in <strong>the</strong> largest size classes than random<br />

sites (Table 4.9). Again, note that <strong>the</strong>se are<br />

relative comparisons, and refer to percentages of<br />

total trees ra<strong>the</strong>r than to absolute numbers of<br />

trees.<br />

Seamans and Gutiérrez (in press) compared<br />

habitat characteristics sampled on 0.04-ha (0.1ac)<br />

circular plots at 78 roost sites and 71 random<br />

sites, Tularosa Mountains, Upper Gila Mountains<br />

RU. Roost plots were centered on <strong>the</strong> roost<br />

tree (one plot each from 78 separate owls), and<br />

random plots were centered on randomlyselected<br />

trees throughout <strong>the</strong> study area. <strong>Owl</strong>s<br />

roosted in mixed-conifer/oak <strong>for</strong>est more than<br />

expected by chance, and in pine-oak <strong>for</strong>est and


23<br />

Table Table 4.9. 4.9. Habitat characteristics sampled on 0.04 ha (0.1 ac) circular plots within home ranges of radio-tagged <strong>Mexican</strong> spotted owls inhabiting<br />

mixed-conifer and ponderosa pine <strong>for</strong>ests, nor<strong>the</strong>rn Arizona (Upper Gila Mountains <strong>Recovery</strong> Unit). Data from Ganey and Balda (1994); n = six<br />

owls occupying four home ranges. Shown are means and standard deviations (in paren<strong>the</strong>ses).<br />

FPO


pinyon-juniper woodland less than expected.<br />

Most roosts were located on <strong>the</strong> lower third on<br />

slopes. Roost sites differed significantly from<br />

random plots <strong>for</strong> several variables (Table 4.10).<br />

Roost sites were best separated from random<br />

sites in a discriminant function analysis by<br />

canopy closure and height variance; both were<br />

greater on roost than on random sites. Crossvalidation<br />

analyses indicated that <strong>the</strong> discriminant<br />

analysis results were stable (Seamans and<br />

Gutiérrez in press).<br />

A. Hodgson and P. Stacey also evaluated<br />

roost sites in <strong>the</strong> Upper Gila Mountains RU<br />

(Peter Stacey, Univ. of Nevada, Reno, NV; pers.<br />

comm.). They compared habitat characteristics<br />

sampled on 0.04-ha (0.1-ac) circular plots<br />

between 55 roost sites and 69 random sites in<br />

<strong>the</strong> San Mateo Mountains, New Mexico (<strong>the</strong><br />

number of plots is also greater than <strong>the</strong> number<br />

of owls in this study). <strong>Owl</strong>s typically roosted in<br />

or near canyon bottoms, in relatively dense<br />

stands of mixed-conifer <strong>for</strong>est containing significantly<br />

more Douglas-fir, Gambel oak, and<br />

limber pine than random sites. Deciduous trees<br />

accounted <strong>for</strong> >28% of total basal area and<br />

>50% of total tree density. Roost trees averaged<br />

31 cm (11.6 in) dbh, with most roosting occurring<br />

in Douglas-fir (54%) or Gambel oak (21%).<br />

In a second comparison, Hodgson and<br />

Stacey restricted <strong>the</strong>ir analysis to roost and<br />

random sites in mixed-conifer <strong>for</strong>est (n = 55 and<br />

36, respectively). Within this <strong>for</strong>est type, roost<br />

Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.10. 4.10. Habitat characteristics at <strong>Mexican</strong> spotted owl roost (n = 78) and random sites in <strong>the</strong><br />

Tularosa Mountains, New Mexico; Upper Gila Mountains RU. Shown are means and standard<br />

deviations (in paren<strong>the</strong>sis). Data from Seamans and Gutiérrez (in press).<br />

FPO<br />

24<br />

sites contained greater densities and basal areas<br />

of Gambel oak and lower densities and basal<br />

areas of conifers than random sites. Differences<br />

between roost and random sites were generally<br />

greatest with respect to trees from 15-30 cm<br />

(5.9-11.8 in) dbh. Roost sites had greater densities<br />

of deciduous trees and lower densities of<br />

coniferous tree in this size-class than random<br />

plots within mixed-conifer <strong>for</strong>est (Peter Stacey,<br />

Univ. of Nevada, Reno, NV; pers. comm.).<br />

Tarango et al. (1994) sampled seven roost<br />

sites in Chihuahua, Mexico (Sierra Madre<br />

Occidental-Norte RU), using 0.04-ha (0.1-ac)<br />

circular plots. Roosts were typically located in<br />

multi-layered pine-oak <strong>for</strong>ests on <strong>the</strong> lower<br />

portions of north-facing slopes. Oaks dominated<br />

most roost sites by density, comprising 46.6% of<br />

<strong>the</strong> trees present, on average (Tarango et al.<br />

1994:1).<br />

Foraging Foraging Sites<br />

Sites<br />

The only available data on <strong>for</strong>aging sites<br />

come from Ganey (1988) and Ganey and Balda<br />

(1994), discussed above. Relative to random<br />

sites, <strong>for</strong>aging sites within owl home ranges had<br />

greater total basal areas and more big down logs/<br />

ha (Table 4.9). Relative to roosting sites, <strong>for</strong>aging<br />

sites had lower canopy closure and fewer<br />

snags/ha (Table 4.9). In a discriminant function<br />

analysis, <strong>for</strong>aging sites were not as readily distinguished<br />

from random sites as roosting sites and


were also more variable along a discriminant<br />

function axis (Ganey 1988, fig. 12). Both of<br />

<strong>the</strong>se results suggest greater variability in <strong>for</strong>aging<br />

habitat than in roosting habitat, consistent<br />

with results at <strong>the</strong> home-range scale (Ganey and<br />

Balda 1994). As noted above <strong>for</strong> roosting sites,<br />

<strong>for</strong>aging sites sampled in this study represented<br />

intensively used habitat and probably do not<br />

represent <strong>the</strong> full range of conditions used <strong>for</strong><br />

<strong>for</strong>aging.<br />

PATTERNS PATTERNS OF OF HABITAT HABITAT USE<br />

USE<br />

AT AT A A TREE TREE SCALE<br />

SCALE<br />

Several studies have examined characteristics<br />

of trees and o<strong>the</strong>r microsites, such as cliff ledges<br />

or caves, used by owls <strong>for</strong> nesting or roosting.<br />

This in<strong>for</strong>mation is primarily descriptive, and<br />

with <strong>the</strong> exception of SWCA (1992), Ruess<br />

(1995), and Seamans and Gutiérrez (in press)<br />

provides no basis <strong>for</strong> comparing used and available<br />

trees.<br />

Nest Nest Trees<br />

Trees<br />

Nest trees were described, with varying levels<br />

of detail, by SWCA (1992), Armstrong et al.<br />

(1994), Fletcher and Hollis (1994), Ruess<br />

(1995), Seamans and Gutiérrez (in press), and<br />

Arizona Game and Fish Department (unpublished<br />

data). All of <strong>the</strong>se studies were conducted<br />

on FS lands in Arizona and New Mexico, and<br />

some nests may be represented in ³2 studies.<br />

SWCA (1992) sampled 84 nest trees; 81%<br />

were conifers and 19% were hardwoods. Fifty<br />

percent of all nests were in Douglas-fir trees,<br />

with 20% and 19% occurring in Gambel oak<br />

and white fir, respectively (SWCA 1992:17).<br />

Eight percent (n = 7) of all nests occurred in<br />

snags; five of <strong>the</strong>se (57%) were Gambel oak<br />

snags (SWCA 1992:17).<br />

Nest trees averaged 63.3 cm (24.9 in) dbh,<br />

and ranged from 17-127 cm (6.7-50.0 in;<br />

SWCA 1992). Nest structures in living oak trees<br />

(n = 14, 17%) were located ei<strong>the</strong>r in a broken<br />

top (n = 3) or a side cavity (n = 11). Nest structures<br />

in live conifers included broken top cavities<br />

(n = 5), old raptor nests (n = 14), witches<br />

brooms (n = 25), stick plat<strong>for</strong>ms on “bayonet<br />

Volume II/Chapter 4<br />

25<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

limbs” (n = 12), stick nests in a multiple-topped<br />

tree (n = 4), and a squirrel nest (n = 1). All snag<br />

nests were ei<strong>the</strong>r broken top (n = 5), or cavity<br />

(n = 3), and one snag contained both nest types<br />

(SWCA 1992:21).<br />

Nest trees were significantly more likely to<br />

have a de<strong>for</strong>med crown than randomly-selected<br />

trees (SWCA 1992:30), although 65% of all nest<br />

trees had a normal crown <strong>for</strong>m. Types of de<strong>for</strong>med<br />

crown included broken top (19%),<br />

multiple top (5%), dead top (10%), and dying<br />

top (1%). These percentages are based on 81<br />

nest trees, with three unaccounted <strong>for</strong> (SWCA<br />

1992: table 9).<br />

Fletcher and Hollis (1994) reported on<br />

microsite characteristics of 248 nests located<br />

during FS inventory and monitoring activities<br />

throughout Arizona and New Mexico. It is<br />

impossible to tell how many different pairs of<br />

owls <strong>the</strong>se nests represent, and some of <strong>the</strong>se<br />

sites may also be included in samples discussed<br />

elsewhere (SWCA 1992, Ruess 1995, Seamans<br />

and Gutiérrez in press). Fur<strong>the</strong>rmore, not all<br />

nests could be assigned to a particular <strong>Recovery</strong><br />

Unit, limiting regional analyses. There<strong>for</strong>e, we<br />

simply summarize some general patterns resulting<br />

from this data set. In many cases, <strong>the</strong> numbers<br />

presented here were recalculated from<br />

summary data in Fletcher and Hollis (1994). In<br />

those cases <strong>the</strong> page or figure number where <strong>the</strong><br />

data were found is cited. Sample sizes vary<br />

among values reported, because not all variables<br />

were sampled at each site.<br />

Of <strong>the</strong> 248 nests sampled, 90.3 and 9.7%<br />

were in trees and cliffs, respectively (Fletcher and<br />

Hollis 1994: fig. 28). Most nests fell within a<br />

fairly narrow elevational band, with 72% falling<br />

between 1982 and 2287 m (6500-7500 ft),<br />

85.4% falling between 1982 and 2591 m (6500-<br />

8500 ft), and 95.5% falling between 1829 and<br />

2591 m (6000-8500 ft), respectively (Fletcher<br />

and Hollis 1994: fig. 29). Forty-three percent of<br />

<strong>the</strong> cliff nests were found below 1982 m (6500<br />

ft).<br />

Almost 50% of 236 nests where aspect was<br />

recorded were located on north or nor<strong>the</strong>ast<br />

aspects (Fletcher and Hollis 1994:48). Slope<br />

averaged 44 ± 40 (SD)%, with 34.5% of all<br />

nests found on slopes >40% (Fletcher and Hollis<br />

1994:49). Almost half of all nests were located


on <strong>the</strong> lower third of slopes, with <strong>the</strong> remainder<br />

split almost evenly between middle and upper<br />

slopes (Fletcher and Hollis 1994: fig. 50).<br />

Roughly 50% of <strong>the</strong> nests on upper slopes were<br />

cliff nests or nests in Gambel oak (Fletcher and<br />

Hollis 1994: fig. 50).<br />

Dominant cover types recorded at 237 nest<br />

sites were: 80.2% mixed-conifer, 15.2% pineoak,<br />

2.1% ponderosa pine, 1.7% riparian, and<br />

0.8% o<strong>the</strong>r (Fletcher and Hollis 1994: fig. 35).<br />

Of 224 tree nests, 57% were in Douglas-fir, 16%<br />

in Gambel oak, 13% in white fir, 9% in ponderosa<br />

pine, and 5% in o<strong>the</strong>r species (Fletcher and<br />

Hollis 1994: fig 40).<br />

Nest tree diameter was recorded at 204 nest<br />

sites. Trees < 15.2 cm (61 cm (>24 in) dbh (Fletcher and Hollis<br />

1994: fig. 41). Forty-five percent of tree nests<br />

were classified as “witches broom”, with 31.3%<br />

in cavities (including broken tops), 14.7% in<br />

“debris plat<strong>for</strong>ms”, and 8.9% in “o<strong>the</strong>r stick<br />

nests” (Fletcher and Hollis 1994: fig. 36).<br />

Six of 11 nests (54.5%) sampled by Ruess<br />

(1995: fig 7) were in Gambel oak, with <strong>the</strong><br />

remainder in ponderosa pine. Ruess (1995) did<br />

not report mean diameters <strong>for</strong> nest trees, but<br />

found four nests (36.4%) in trees of<br />

presettlement origin (>115 yrs in age) despite<br />

<strong>the</strong> fact that such trees accounted <strong>for</strong> only 0.5%<br />

of total trees on his study area (Ruess 1995: table<br />

3, fig. 7).<br />

Seamans and Gutiérrez (in press) reported<br />

78% of 27 nests in Douglas-fir, 11% in white fir,<br />

7% in ponderosa pine, and 4% in southwestern<br />

white pine. With respect to nest structure, 61%<br />

were located in dwarf mistletoe infections,<br />

10.5% in old squirrel nests, 10.5% in old raptor<br />

nests, 7% in debris collections, 7% in tree<br />

cavities, and one nest (4%, n = 28 nests <strong>for</strong> <strong>the</strong>se<br />

calculations) was on a cliff. Nest trees averaged<br />

60.6 ± 22.4 cm (23.9 ± 17.6 in) in dbh and<br />

164 ± 44.8 years in age. Nest trees were significantly<br />

larger and older than randomly sampled<br />

trees within <strong>the</strong> nest vicinity (Seamans and<br />

Gutiérrez in press).<br />

Volume II/Chapter 4<br />

26<br />

Roost Roost Trees<br />

Trees<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Several researchers have described characteristics<br />

of roost trees and a very small area around<br />

<strong>the</strong>m. Results are summarized by study area and<br />

RU, where possible, in Table 4.11. In general,<br />

roost characteristics appeared to be relatively<br />

variable among study areas (Table 4.11), suggesting<br />

that greater variability exists among roost<br />

trees than among trees used <strong>for</strong> nesting. Canopy<br />

closure was more consistent among study areas<br />

than most o<strong>the</strong>r characteristics sampled, and was<br />

relatively high on most study areas. Canopy<br />

closure was 85% of<br />

<strong>the</strong> roost sites sampled on that study area were<br />

located on cliffs or in pinyon-juniper woodland<br />

(Table 4.11a).<br />

Roost tree characteristics appeared to be<br />

relatively similar among similar habitat types.<br />

For example, mean roost tree diameter was more<br />

consistent among <strong>the</strong> mesic mixed-conifer sites<br />

(San Francisco Peaks, White Mountains, and<br />

Sacramento Mountains: mixed-conifer) than<br />

between <strong>the</strong>se sites and o<strong>the</strong>r areas. Similarly,<br />

characteristics were more similar among <strong>the</strong><br />

more xeric sites (Sacramento Mountains: xeric<br />

mixed <strong>for</strong>est and Bar-M watershed) than between<br />

<strong>the</strong>se and o<strong>the</strong>r areas (Table 4.11). One<br />

clear pattern that emerges from <strong>the</strong>se studies is<br />

that owls roost in smaller trees, on average, than<br />

those used <strong>for</strong> nesting. Ruess (1995: fig. 7),<br />

however, noted that three of nine roosts observed<br />

in ponderosa pine-Gambel oak <strong>for</strong>est<br />

were >115 yrs old, despite <strong>the</strong> fact that only<br />

0.5% of all trees sampled in <strong>the</strong> area were of<br />

such age (Ruess 1995: Table 3). This suggests<br />

that old, large trees may also be important <strong>for</strong><br />

roosting in some areas.<br />

Fletcher and Hollis (1994) reported summary<br />

characteristics sampled at 433 roost sites<br />

located during FS inventory and monitoring<br />

ef<strong>for</strong>ts in Arizona and New Mexico. This sample<br />

is discussed separately here because roost sites<br />

could not generally be assigned to particular<br />

RUs, as was done <strong>for</strong> <strong>the</strong> data sets summarized in<br />

Table 4.11. This data set is subject to all <strong>the</strong>


Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.11a. 4.11a. Selected characteristics of roost sites used by radio-tagged Mexcan spotted owls in <strong>the</strong><br />

Colorado Plateau <strong>Recovery</strong> Unit. Values shown are means (± standard deviation) <strong>for</strong> continuous<br />

variables, % <strong>for</strong> categorical variables. Source: D.W. Willey (unpublished data).<br />

FPO<br />

27


Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.11b. 4.11b. Selected characteristics of roost sites used by radio-tagged <strong>Mexican</strong> spotted owls in <strong>the</strong><br />

Upper Gila Mountains <strong>Recovery</strong> Unit. Values shown are mean (± standard deviation) <strong>for</strong> continuous<br />

variables, % <strong>for</strong> categorical variables.<br />

FPO<br />

28


Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.11c. 4.11c. Selected characteristics of roost sites used by radio-tagged <strong>Mexican</strong> spotted owls in <strong>the</strong><br />

Basin and East-Range <strong>Recovery</strong> Unit. Values shown are mean (± standard deviation) <strong>for</strong> continuous<br />

variables, % <strong>for</strong> categorical variables.<br />

FPO<br />

29


limitations discussed relative to nest sites (see<br />

above) described by Fletcher and Hollis (1994).<br />

Ninety-five percent of all roosts were in<br />

trees, with <strong>the</strong> remainder on cliffs or in caves<br />

(Fletcher and Hollis 1994: fig. 42). Percent slope<br />

averaged 46 ± 34%, with 45% of all roost sites<br />

occurring on slopes >40% (Fletcher and Hollis<br />

1994: fig. 46). Most (57%) of <strong>the</strong>se roost sites<br />

were found on <strong>the</strong> lower third of slopes (Fletcher<br />

and Hollis 1994: fig 47).<br />

Of 367 roost sites where cover type was<br />

recorded, 65.7% were in mixed-conifer, 27.2%<br />

in pine-oak, 2.7% in riparian, 1.4% in ponderosa<br />

pine, and 3.0% in o<strong>the</strong>r cover types<br />

(Fletcher and Hollis 1994: fig. 49). A variety of<br />

trees provided roost perches, including Douglasfir<br />

(38.3%), Gambel oak (17.9%), ponderosa<br />

pine (11.9%), white fir (11.4%), o<strong>the</strong>r oak<br />

species (8.0%), and o<strong>the</strong>r species (12.4%;<br />

Fletcher and Hollis 1994: fig. 53; n = 402). Trees<br />

24 in) in dbh, respectively<br />

(Fletcher and Hollis 1994: fig. 54).<br />

Ganey and Block (unpublished data) evaluated<br />

seasonal differences in roost site characteristics<br />

in ponderosa pine-Gambel oak <strong>for</strong>est (Table<br />

4.12). Canopy closure was greater at roosts used<br />

during <strong>the</strong> breeding season. <strong>Owl</strong>s used Gambel<br />

oak significantly more often during <strong>the</strong> breeding<br />

season, and tended to roost more often on <strong>the</strong><br />

upper third of slopes during <strong>the</strong> breeding season<br />

and on <strong>the</strong> middle third during <strong>the</strong> nonbreeding<br />

season.<br />

Ganey and Block (unpublished data) also<br />

reported some characteristics of winter roosts<br />

used by two owls that migrated to lower elevations<br />

(Table 4.13). Both owls roosted primarily<br />

in pinyon-juniper woodland, on <strong>the</strong> middle and<br />

upper portions of slopes. They typically perched<br />

low in short juniper trees, well hidden by dense<br />

foliage and near <strong>the</strong> center of <strong>the</strong> tree. These<br />

winter roosts differed greatly in structure and<br />

species composition from typical summer<br />

roosting habitat.<br />

Volume II/Chapter 4<br />

30<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

WINTERING WINTERING HABITAT HABITAT<br />

HABITAT<br />

Present knowledge of wintering habitat of<br />

<strong>Mexican</strong> spotted owls comes primarily from<br />

radiotelemetry studies (see Patterns of Habitat<br />

Use at <strong>the</strong> Home Range Scale) and opportunistic<br />

observations of wintering adults. Radiotelemetry<br />

studies indicate that many owls remain on <strong>the</strong>ir<br />

breeding areas throughout <strong>the</strong> year, whereas<br />

some migrate off of <strong>the</strong> study area. Where<br />

wintering areas of migrants have been located,<br />

<strong>the</strong>y are typically in lower elevation woodland or<br />

scrub habitats with more open structure than<br />

typical breeding habitat. However, one owl in<br />

<strong>the</strong> Colorado Plateau RU migrated upwards in<br />

elevation to winter in coniferous <strong>for</strong>est (Willey<br />

1993).<br />

Opportunistic sightings of spotted owls<br />

during <strong>the</strong> winter also suggest that part of <strong>the</strong><br />

population moves to lower elevations. For<br />

example, owls have been sighted in lower Sabino<br />

Canyon, outside of Tucson, Arizona, and on golf<br />

courses in Tucson in recent winters (Russell<br />

Duncan, Southwestern Field Biologists, Tucson,<br />

AZ, pers. comm.). An adult owl banded on <strong>the</strong><br />

Gila National Forest, Upper Gila Mountains<br />

RU, was recovered during winter 1995 near<br />

Deming, New Mexico, Basin and Range-East<br />

RU. This bird had apparently traveled approximately<br />

160 km (100 mi) from <strong>the</strong> area where it<br />

was located during <strong>the</strong> breeding season, from<br />

high-elevation <strong>for</strong>est to Chihuahuan desert<br />

(Mark Seamans, Humboldt State Univ., Arcata,<br />

CA, pers. comm.).<br />

In summary, available evidence on wintering<br />

habitat is limited, but suggests that <strong>the</strong> bulk of<br />

<strong>the</strong> owl population is nonmigratory. Where<br />

migration does occur, it typically involves<br />

movement to lower, warmer, and more open<br />

habitats. In some cases, migration involves<br />

movement between adjacent <strong>Recovery</strong> Units.<br />

Little quantitative data exists to describe typical<br />

wintering habitat <strong>for</strong> ei<strong>the</strong>r migrants or yearround<br />

residents.<br />

DISPERSAL DISPERSAL HABITAT<br />

HABITAT<br />

Very little is known about habitat use ei<strong>the</strong>r<br />

by adults during migration or by juveniles


Volume II/Chapter 4<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.12. 4.12. Seasonal roost site characteristics of radio-tagged <strong>Mexican</strong> spotted owls in ponderosa<br />

pine-Gambel oak <strong>for</strong>est, Arizona (Upper Gila Mountains <strong>Recovery</strong> Unit). Data from Ganey and<br />

Block (unpublished). Values shown are mean (± SD) <strong>for</strong> continuous variables, % <strong>for</strong> categorical variables.<br />

FPO<br />

31


during dispersal. Willey (1993) monitored seven<br />

dispersing juvenile owls in Utah. These juveniles<br />

apparently moved through a variety of habitat<br />

types, including several that might generally be<br />

considered too open <strong>for</strong> use by spotted owls.<br />

None of <strong>the</strong>se juveniles survived to reproduce.<br />

A. Hodgson and P. Stacey radio tagged five<br />

juveniles in <strong>the</strong> San Mateo Mountains, New<br />

Mexico (Upper Gila Mountains RU). Two<br />

juveniles apparently dispersed across open<br />

grassland to <strong>the</strong> Black Range, but <strong>the</strong>ir ultimate<br />

fate is not known (Peter Stacey, Univ. of Nevada,<br />

Reno, pers. comm.). Three o<strong>the</strong>r juveniles<br />

apparently remained in <strong>the</strong> San Mateo Mountains.<br />

No in<strong>for</strong>mation is available on habitats<br />

used by <strong>the</strong>se owls.<br />

Volume II/Chapter 4<br />

ADDITIONAL ADDITIONAL STUDIES<br />

STUDIES<br />

In addition to <strong>the</strong> above studies, two additional<br />

studies are treated separately here because<br />

<strong>the</strong>y ei<strong>the</strong>r were not specific to a single scale<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 4.13. 4.13. 4.13. Characteristics of roost sites used by two migrant <strong>Mexican</strong> spotted owls on <strong>the</strong>ir winter<br />

range. Both owls bred in ponderosa pine-Gambel oak <strong>for</strong>est in <strong>the</strong> Upper Gila Mountains <strong>Recovery</strong><br />

Unit, and wintered in pinyon-juniper woodland on <strong>the</strong> Basin and Range-West <strong>Recovery</strong> Unit. Data<br />

from Ganey and Block (unpublished).<br />

FPO<br />

32<br />

(Johnson 1989, Dames and Moore 1990) or did<br />

not distinguish between site types (Dames and<br />

Moore 1990). Johnson (1989) compared habitat<br />

characteristics sampled at one roost and one nest<br />

site with characteristics sampled at 20 points<br />

within <strong>the</strong> same stand. Thus, this provides a<br />

comparison of site characteristics with overall<br />

characteristics of <strong>the</strong> surrounding stand. Mean<br />

tree height, overstory basal area, understory basal<br />

area, overstory density, and snag density were all<br />

greater at <strong>the</strong> roost and nest sites than in <strong>the</strong><br />

surrounding stand. Despite <strong>the</strong> greater understory<br />

basal area at roost and nest sites, understory<br />

density was greater within <strong>the</strong> stand, suggesting<br />

that <strong>the</strong> understory at <strong>the</strong> roost and nest sites<br />

contained fewer but larger trees (Johnson<br />

1989:12). Johnson (1989:14-15) fur<strong>the</strong>r noted<br />

that <strong>the</strong> roost and nest sites ranked high (relative<br />

to <strong>the</strong> stand) <strong>for</strong> evenness indices <strong>for</strong> both tree<br />

species and diameter, and suggested that smallscale<br />

diversity may be an important factor in


habitat selection by spotted owls (see also<br />

Johnson and Johnson 1988, Johnson 1990).<br />

Dames and Moore (1990) reported on<br />

habitat characteristics in areas occupied by<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong>s in Arizona and New<br />

Mexico (Upper Gila Mountains and Basin and<br />

Range East-RUs). Sampling methods and<br />

sampled area differed between states (Dames and<br />

Moore 1990:10), and most of <strong>the</strong> area sampled<br />

was based on owl presence in <strong>the</strong> area, ra<strong>the</strong>r<br />

than on specific evidence of use by owls <strong>for</strong><br />

ei<strong>the</strong>r <strong>for</strong>aging, roosting, or nesting. Results of<br />

hypo<strong>the</strong>sis tests regarding habitat characteristics<br />

in this study were inconclusive. In both states,<br />

<strong>the</strong> most consistent feature within and among<br />

areas sampled was variability (Dames and Moore<br />

1990: executive summary).<br />

Volume II/Chapter 4<br />

CONCLUSIONS<br />

CONCLUSIONS<br />

Habitat Habitat Relationships<br />

Relationships<br />

Several patterns are evident upon evaluation<br />

of current knowledge regarding habitat relationships<br />

of <strong>Mexican</strong> spotted owls. The first is that<br />

most in<strong>for</strong>mation is limited to relatively fine<br />

spatial scales. For example, we have considerable<br />

in<strong>for</strong>mation about roost and nest trees, and roost<br />

and nest sites, but have little in<strong>for</strong>mation on<br />

stands used by owls, habitat composition of owl<br />

home ranges, or landscape configurations used<br />

by spotted owls. Second, most in<strong>for</strong>mation on<br />

owl habitat use relates to <strong>the</strong> breeding season,<br />

and to habitats used <strong>for</strong> nesting and/or roosting.<br />

We know little about habitat use during winter<br />

or dispersal periods, or about what constitutes<br />

adequate <strong>for</strong>aging habitat. Third, most in<strong>for</strong>mation<br />

on owl habitat use and selection comes<br />

from correlative studies that do not demonstrate<br />

cause-and-effect relationships. Fourth, our<br />

knowledge of owl habitat-use patterns comes<br />

from a very short time period (mainly 1984present).<br />

All of <strong>the</strong>se factors limit our ability to define<br />

what constitutes spotted owl habitat, or what<br />

desired future conditions should be <strong>for</strong> spotted<br />

owls. What we can do is describe features of<br />

breeding-season roosting and nesting habitat<br />

used by owls at this time. In most cases, histori-<br />

33<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

cal in<strong>for</strong>mation is not adequate to know whe<strong>the</strong>r<br />

currently-occupied sites were also occupied in<br />

<strong>the</strong> past, or to allow us to draw conclusions<br />

about structural features of habitats used in <strong>the</strong><br />

past.<br />

<strong>Spotted</strong> owls typically nest or roost ei<strong>the</strong>r in<br />

deep, rocky canyons, or in any of several <strong>for</strong>est<br />

cover types. The relative use of canyons versus<br />

<strong>for</strong>ests varies among regions. For example, owls<br />

in parts of <strong>the</strong> Colorado Plateau RU are found<br />

exclusively in deep, rocky canyons, whereas owls<br />

in many o<strong>the</strong>r RUs are found primarily in <strong>for</strong>ests<br />

(although <strong>the</strong>se <strong>for</strong>ests are often in canyons).<br />

Where owls occur in <strong>for</strong>ests, <strong>the</strong>y typically<br />

select large, old trees <strong>for</strong> nesting. Nest and roost<br />

sites are found primarily in mixed-conifer <strong>for</strong>est<br />

(Table 4.1), although pine-oak <strong>for</strong>ests are also<br />

used in some areas, such as parts of <strong>the</strong> Upper<br />

Gila Mountains, Basin and Range-West, and<br />

Sierra Madre Occidental-Norte RUs (Table 4.1;<br />

see also Ganey and Balda 1989a, Duncan and<br />

Taiz 1992, Ganey et al. 1992, Fletcher and<br />

Hollis 1994, Tarango et al. 1994, Seamans and<br />

Gutiérrez in press). Nest and roost sites typically<br />

contain structurally-complex, uneven-aged<br />

<strong>for</strong>ests, with a variety of age- and/or size- classes,<br />

a large tree component, many snags and down<br />

logs, and relatively high basal area and canopy<br />

closure (Tables 4.6-4.11; see also SWCA 1992,<br />

Armstrong et al. 1994, Ganey and Balda 1994,<br />

Ruess 1995, Seamans and Gutiérrez in press).<br />

Diversity of tree species and diameters appears to<br />

be high in many owl nesting and roosting areas<br />

(Johnson and Johnson 1988, Johnson 1989,<br />

Johnson 1990, Seamans and Gutiérrez in press).<br />

Many roost and nest sites are found in canyon<br />

bottoms or low on canyon slopes (Table 4.11;<br />

see also Ganey and Balda 1989a, Fletcher and<br />

Hollis 1994, Tarango et al. 1994:36, Seamans<br />

and Gutiérrez in press). Many have a conspicuous<br />

broadleaved component, in <strong>the</strong> <strong>for</strong>m of<br />

riparian trees or especially various oaks (Ganey<br />

and Balda 1989a, Duncan and Taiz 1992, Ganey<br />

et al. 1992, SWCA 1992, Tarango et al. 1994,<br />

Ruess 1995, Seamans and Gutiérrez in press).<br />

The reasons why spotted owls nest and roost<br />

in structurally-complex, diverse <strong>for</strong>ests and deep<br />

canyons have not been conclusively demonstrated.<br />

Barrows (1981) suggested that owls seek<br />

dense <strong>for</strong>est stands as protection from high


daytime temperatures. This explanation is<br />

attractive with respect to <strong>Mexican</strong> spotted owls,<br />

because <strong>the</strong> most apparent common denominator<br />

between <strong>the</strong> types of <strong>for</strong>ests and deep, rocky<br />

canyons used is that both situations provide cool<br />

microsites (Kertell 1977, Ganey et al. 1988,<br />

Ganey and Balda 1989a, Rinkevich 1991, Willey<br />

1993). Fur<strong>the</strong>r, <strong>the</strong>re is some evidence that,<br />

relative to <strong>the</strong> great horned owl, which is found<br />

in hotter and drier areas, <strong>the</strong> spotted owl has<br />

difficulty dissipating metabolic heat at high<br />

temperatures (Ganey et al. 1993).<br />

Carey (1985) and Gutiérrez (1985) also<br />

hypo<strong>the</strong>sized that nor<strong>the</strong>rn spotted owls might<br />

seek dense old, closed-canopy <strong>for</strong>ests because<br />

such areas supported higher prey densities, or<br />

because <strong>the</strong> owls were better able to avoid<br />

predators in such areas. Present in<strong>for</strong>mation,<br />

however, suggests that owls <strong>for</strong>age in a wider<br />

variety of <strong>for</strong>est types than are used <strong>for</strong> roosting<br />

(Ganey and Balda 1994, see also Ward and Block<br />

1995). Fur<strong>the</strong>r, although owls in one study<br />

roosted primarily in unlogged mixed-conifer<br />

<strong>for</strong>ests, <strong>the</strong>y did not show a strong pattern of<br />

selection <strong>for</strong> such <strong>for</strong>ests when <strong>for</strong>aging (Table<br />

4.4). This suggests that <strong>the</strong> association between<br />

spotted owls and mixed-conifer <strong>for</strong>est may be<br />

driven more by roosting and/or nesting behavior<br />

than by <strong>for</strong>aging behavior (Ganey and Balda<br />

1994). We currently have no in<strong>for</strong>mation with<br />

which to test <strong>the</strong> hypo<strong>the</strong>sis that spotted owls<br />

are better able to avoid predators in complex<br />

<strong>for</strong>ests or deep, rocky canyons.<br />

In summary, at present we suspect that<br />

selection of typical nesting and roosting habitat<br />

is driven primarily by microclimatic considerations.<br />

Prey availability may also be an important<br />

consideration, however, and prey density in and<br />

around an area with microclimatic conditions<br />

typical of nest/roost habitat may determine<br />

whe<strong>the</strong>r or not that site is used by owls. Prey<br />

availability may also determine how large an area<br />

owls must use to meet <strong>the</strong>ir energetic needs<br />

(Carey et al. 1992, Verner et al. 1992, Zabel et<br />

al. 1995). However, we suspect that <strong>the</strong> primary<br />

factor limiting spotted owl distribution is <strong>the</strong><br />

presence on <strong>the</strong> landscape of habitat suitable <strong>for</strong><br />

roosting and nesting.<br />

In some areas, <strong>the</strong> types of <strong>for</strong>ests used <strong>for</strong><br />

roosting and nesting are primarily restricted to<br />

Volume II/Chapter 4<br />

34<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

canyon situations. This is particularly true at<br />

lower elevations, where mixed-conifer <strong>for</strong>est is<br />

generally found only in canyon bottoms or on<br />

north-facing canyon slopes. Thus, <strong>the</strong> distribution<br />

of roosting and nesting habitat, and of<br />

spotted owls in <strong>the</strong>se areas, is naturally fragmented<br />

and discontinuous. Opportunities <strong>for</strong><br />

increasing <strong>the</strong> amount of spotted owl habitat in<br />

such areas are limited, and management ef<strong>for</strong>ts<br />

would be better focused on preserving and<br />

enhancing habitat where it exists. Replacement<br />

habitat in such areas may need to develop in situ<br />

following stand-replacing disturbances.<br />

In o<strong>the</strong>r areas, such as high-elevation mixedconifer<br />

<strong>for</strong>est, it may be possible to develop<br />

spotted owl habitat over more of <strong>the</strong> landscape.<br />

An example of such an area is <strong>the</strong> Sacramento<br />

Mountains (Lincoln National Forest, Basin and<br />

Range-East RU). <strong>Spotted</strong> owls are abundant and<br />

widely distributed in mixed-conifer <strong>for</strong>ests across<br />

this range (Skaggs and Raitt 1988, Fletcher and<br />

Hollis 1994: fig. 24). Much of this area was<br />

subject to relatively intensive railroad logging<br />

early this century (Glover 1984), but <strong>the</strong>se<br />

<strong>for</strong>ests have recovered quickly and attained<br />

structural complexity (Table 4.8b), demonstrating<br />

that development of replacement habitat is<br />

possible after management under some circumstances.<br />

In areas such as this, managers might<br />

combine protection of existing habitat with<br />

attempts to develop replacement habitat over<br />

time, in a more dynamic approach to habitat<br />

management.<br />

Comparisons Comparisons With With O<strong>the</strong>r O<strong>the</strong>r Subspecies<br />

Subspecies<br />

of of of <strong>Spotted</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong>s<br />

<strong>Owl</strong>s<br />

There are many similarities in habitat-use<br />

patterns of <strong>the</strong> three subspecies (nor<strong>the</strong>rn,<br />

Cali<strong>for</strong>nia, and <strong>Mexican</strong>) of spotted owls. For<br />

example, all three appear to be most common in<br />

structurally-complex <strong>for</strong>est environments,<br />

although floristic composition of habitats used<br />

varies both within and between subspecies’<br />

ranges (<strong>for</strong> o<strong>the</strong>r subspecies see reviews in<br />

Thomas et al. 1990, Gutiérrez et al. 1992). Both<br />

<strong>the</strong> Cali<strong>for</strong>nia (Gutiérrez et al. 1992) and <strong>Mexican</strong><br />

subspecies most commonly nest in mixedconifer<br />

<strong>for</strong>est, followed by <strong>for</strong>est types domi-


nated by oaks or conifers and oaks. Both <strong>the</strong><br />

Cali<strong>for</strong>nia (Gutiérrez et al. 1992) and <strong>Mexican</strong><br />

subspecies appear to use a wider variety of<br />

habitat conditions <strong>for</strong> <strong>for</strong>aging than <strong>for</strong> roosting<br />

and nesting. These consistencies in habitat use<br />

patterns between subspecies occupying different<br />

geographic areas and habitat types fur<strong>the</strong>r<br />

streng<strong>the</strong>n our conclusion that spotted owls are<br />

seeking particular types of habitat features <strong>for</strong><br />

nesting and roosting.<br />

Habitat Habitat Trends<br />

Trends<br />

Because <strong>the</strong> listing of <strong>the</strong> owl was based<br />

partially on projected declines in owl habitat, it<br />

is worth discussing what we know about trends<br />

in owl habitat. The following discussion focuses<br />

separately on various habitat conditions used by<br />

spotted owls. It is largely restricted to trends in<br />

roosting and nesting habitat, because most of<br />

our in<strong>for</strong>mation relates to such areas, and because<br />

such habitat is thought to limit spotted<br />

owl distribution. Finally, because little historical<br />

in<strong>for</strong>mation exists with respect to <strong>the</strong> type of<br />

microsite conditions that describe roosting and<br />

nesting habitat of spotted owls, this discussion is<br />

necessarily largely qualitative.<br />

Rocky Rocky Canyons<br />

Canyons<br />

<strong>Mexican</strong> spotted owls are found primarily in<br />

rocky canyons in parts of <strong>the</strong>ir range, such as<br />

sou<strong>the</strong>rn Utah (Kertell 1977, Rinkevich 1991,<br />

Willey 1992, 1993, Utah <strong>Mexican</strong> spotted owl<br />

Technical Team 1994). In o<strong>the</strong>r areas, part of <strong>the</strong><br />

population also inhabits rocky canyons, but<br />

<strong>the</strong>se are generally more heavily <strong>for</strong>ested than<br />

<strong>the</strong> slickrock canyons found in parts of <strong>the</strong><br />

Colorado Plateau RU (Ganey and Balda 1989a,<br />

USDA Forest Service 1993, Fletcher and Hollis<br />

1994). There is little evidence <strong>for</strong> change in<br />

habitat quality or loss of habitat in slickrock<br />

canyons. Canyon-bottom vegetation may have<br />

been degraded by grazing in some cases (see<br />

below), and some habitat may have been lost<br />

beneath large reservoirs along <strong>the</strong> Colorado<br />

River and its tributaries. The latter change could<br />

have decreased connectivity among remaining<br />

populations. O<strong>the</strong>rwise, we suspect that habitat<br />

Volume II/Chapter 4<br />

35<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

trends are relatively stable in areas where owls are<br />

found primarily in slickrock canyons.<br />

Riparian Riparian Forests<br />

Forests<br />

Historically, owls were found in low-elevation<br />

riparian <strong>for</strong>ests (Bendire 1892, Phillips et al.<br />

1964, and possibly Woodhouse 1853). These<br />

<strong>for</strong>ests have undergone extensive modification<br />

because of recreation, flood control, livestock<br />

grazing, and modification of natural water tables<br />

(Knopf et al. 1988; see also Kennedy 1977,<br />

Kauffman and Krueger 1984, Minckley and<br />

Clark 1984, Skovlin 1984, Minckley and Rinne<br />

1985, Platts 1990, Schulz and Leininger 1990,<br />

Dick-Peddie 1993). Collectively, <strong>the</strong>se activities<br />

and o<strong>the</strong>r factors have changed <strong>the</strong> species<br />

composition and structure of riparian <strong>for</strong>ests. In<br />

extreme cases, riparian <strong>for</strong>ests have disappeared<br />

entirely. No breeding spotted owls have been<br />

documented in lowland riparian <strong>for</strong>ests in recent<br />

times. Surveys of such habitat have been far from<br />

exhaustive, and owls may still inhabit some<br />

remnant riparian <strong>for</strong>ests. Never<strong>the</strong>less, <strong>the</strong><br />

overall trend with respect to spotted owls breeding<br />

in lowland riparian <strong>for</strong>est habitat has clearly<br />

been negative.<br />

<strong>Spotted</strong> owls also commonly occur in<br />

canyon-bottom riparian <strong>for</strong>ests at higher elevations,<br />

interspersed with o<strong>the</strong>r <strong>for</strong>est types. In<br />

many cases <strong>the</strong>se <strong>for</strong>ests have also been degraded<br />

(Schulz and Leininger 1991, see also Fleischner<br />

1994). Management to retain and enhance <strong>the</strong>se<br />

riparian <strong>for</strong>ests would likely be beneficial to<br />

spotted owls (as well as many o<strong>the</strong>r plants and<br />

animals).<br />

Coniferous Coniferous Coniferous Forests<br />

Forests<br />

Trends in coniferous <strong>for</strong>ests are more difficult<br />

to evaluate. It is clear that changes have<br />

occurred in southwestern <strong>for</strong>ests. These stem<br />

primarily from three sources: disruption of<br />

natural disturbance regimes, grazing, and timber<br />

harvest. Many decades of fire suppression have<br />

disrupted natural disturbance regimes (Cooper<br />

1960, Madany and West 1983, Stein 1988,<br />

Savage and Swetnam 1990, Covington and<br />

Moore 1992, 1994, Harrington and Sackett<br />

1992, Johnson 1995). The absence of frequent,


low-intensity fire, coupled with widespread<br />

overgrazing by livestock in <strong>the</strong> late 1800s,<br />

reduced competition between herbaceous<br />

vegetation and tree seedlings. These effects<br />

produced a good seedbed <strong>for</strong> conifer regeneration,<br />

and generally resulted in increases in tree<br />

densities on <strong>for</strong>ested lands (Rummel 1951,<br />

Madany and West 1983, Zimmerman and<br />

Neuenschwander 1984, Stein 1988, Savage and<br />

Swetnam 1990, Covington and Moore 1992,<br />

1994, Harrington and Sackett 1992, Johnson<br />

1995, Ruess 1995). Such increases in tree density<br />

can not only alter stand structure, but can also<br />

lead to declines in shade-intolerant tree species,<br />

and drive ecological succession from one <strong>for</strong>est<br />

type to ano<strong>the</strong>r. For example, some ponderosa<br />

pine <strong>for</strong>ests appear to be converting to mixedconifer<br />

<strong>for</strong>est.<br />

Timber harvest in many areas has also altered<br />

stand structure and sometimes species composition.<br />

Timber harvest can occur in different<br />

<strong>for</strong>ms, with different intensity, and with different<br />

effects on stand structure. In many cases,<br />

however, <strong>the</strong> net effect of timber harvest is a<br />

decrease in old trees and at least a short-term<br />

decrease in tree density and basal area. In <strong>the</strong>se<br />

respects timber harvest tends to work opposite<br />

<strong>the</strong> effects of disruptions in natural disturbance<br />

regimes described above. Where timber harvest<br />

targets shade-intolerant species, however, it<br />

could fur<strong>the</strong>r <strong>the</strong> trend of ecological succession<br />

towards shade-tolerant species.<br />

The net effect of <strong>the</strong>se processes on amount<br />

and quality of spotted owl habitat is difficult to<br />

determine. USDI (1993:14251), citing Fletcher<br />

(1990), postulated that spotted owl habitat had<br />

decreased in amount due to timber harvest.<br />

Fletcher and Hollis (1994:29) estimated that<br />

1,068,500 ac of <strong>for</strong>ested habitat in Arizona and<br />

New Mexico had been rendered unsuitable <strong>for</strong><br />

spotted owls due to human activities, primarily<br />

timber harvest, through 1993. Conversely, Hull<br />

(1995:7) reported that “thousands of acres have<br />

converted from ponderosa pine to mixedconifer,”<br />

and argued that <strong>the</strong> amount of <strong>Mexican</strong><br />

spotted owl habitat had increased. For several<br />

reasons, we submit that ei<strong>the</strong>r viewpoint is<br />

difficult to substantiate with presently-available<br />

data.<br />

Volume II/Chapter 4<br />

36<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

First, <strong>the</strong> claim that thousands of acres of<br />

ponderosa pine have converted to mixed-conifer<br />

is unsubstantiated. Johnson (1995: fig.1) shows a<br />

large increase in acreage of mixed-conifer in<br />

Arizona and New Mexico between 1966 and<br />

1986, based on <strong>for</strong>est inventories conducted in<br />

those years. Johnson (1995:1) also presents<br />

numbers suggesting that much of <strong>the</strong> increase in<br />

mixed-conifer <strong>for</strong>est is due to invasion of meadows,<br />

ra<strong>the</strong>r than conversion of ponderosa pine<br />

<strong>for</strong>est. Fur<strong>the</strong>r, this in<strong>for</strong>mation is extrapolated<br />

from data presented in <strong>the</strong> original inventories<br />

(Choate 1966, Spencer 1966, Conner et al.<br />

1990, Van Hooser et al. 1993), and no in<strong>for</strong>mation<br />

is provided on how that extrapolation was<br />

done. There<strong>for</strong>e, it is impossible to assess <strong>the</strong><br />

accuracy of <strong>the</strong> figures presented. Methods and<br />

definitions may also have changed between<br />

inventories. Referring to comparisons between<br />

<strong>the</strong> 1986 inventory and earlier inventories, Van<br />

Hooser et al. (1993:1) state: “The changes in<br />

definitions and survey standards make detailed<br />

comparisons with previous inventory results<br />

unwise.”<br />

Second, <strong>the</strong> idea that all mixed-conifer <strong>for</strong>est<br />

is spotted owl habitat is not supported by <strong>the</strong><br />

available data. As noted previously, owls roost<br />

and nest primarily in structurally-complex<br />

<strong>for</strong>ests with particular features, including large,<br />

old trees. Recent invasion of meadows by mixedconifer<br />

<strong>for</strong>est is unlikely to have created this type<br />

of habitat. If that process proceeds, however, it<br />

could create owl habitat in <strong>the</strong> future.<br />

Finally, most available data on historical<br />

<strong>for</strong>est structure and increases in <strong>for</strong>est density<br />

relate to ponderosa pine <strong>for</strong>est (USGS 1904,<br />

Woolsey 1911, Harrington and Sackett 1984,<br />

Covington and Moore 1992, 1994). As we have<br />

shown here, this <strong>for</strong>est type is not typically used<br />

<strong>for</strong> roosting and nesting by spotted owls. Data<br />

presented by Ruess (1995) suggest that similar<br />

changes have occurred in ponderosa pine-<br />

Gambel oak <strong>for</strong>est, which is used by spotted<br />

owls. No such data exist <strong>for</strong> mixed-conifer <strong>for</strong>est,<br />

however, which is <strong>the</strong> primary type used by<br />

spotted owls <strong>for</strong> roosting and especially <strong>for</strong><br />

nesting. It seems logical to assume that increases<br />

in density have also occurred within this <strong>for</strong>est<br />

type. Without quantitative data on changes in<br />

this <strong>for</strong>est type, however, determining whe<strong>the</strong>r


or not such changes have been favorable to<br />

spotted owls is essentially impossible.<br />

In summary, conflicting speculation exists<br />

regarding trends in spotted owl habitat in<br />

coniferous <strong>for</strong>est. Some authors speculate that<br />

timber harvest has reduced <strong>the</strong> amount of owl<br />

habitat, whereas o<strong>the</strong>rs speculate that fire suppression<br />

has increased <strong>the</strong> amount of owl habitat.<br />

Data presented here suggest that spotted owl<br />

habitat is complex. The types of historical<br />

evidence available do not allow <strong>for</strong> any clear<br />

analysis of trends in such habitat. We recognize<br />

that we cannot return to <strong>the</strong> past to collect such<br />

data, but we can learn from this situation. Our<br />

inability to evaluate habitat trends strongly<br />

emphasizes <strong>the</strong> need <strong>for</strong> accurate and comprehensive<br />

inventory and monitoring of <strong>for</strong>est<br />

resources, so that in <strong>the</strong> future changes in <strong>for</strong>est<br />

habitat can be assessed over time.<br />

Research Research Research Considerations<br />

Considerations<br />

Clearly, much remains to be learned about<br />

habitat relationships of <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Gutiérrez et al. (1992) outlined a number of<br />

research considerations relative to habitat relationships<br />

of <strong>the</strong> Cali<strong>for</strong>nia spotted owl. These<br />

are all relevant to <strong>the</strong> <strong>Mexican</strong> subspecies as well.<br />

We briefly summarize some additional considerations<br />

here.<br />

Quantitative data on patterns of owl habitat<br />

use are largely or completely lacking <strong>for</strong> some<br />

geographic regions, habitat types, and spatial<br />

scales. Particularly striking is <strong>the</strong> lack of quantitative<br />

data on habitat relationships at <strong>the</strong> stand<br />

and landscape scales. Some estimates of habitat<br />

conditions measured on small plots, such as<br />

canopy closure, may not be representative of<br />

conditions at <strong>the</strong> stand scale. Fur<strong>the</strong>r, many<br />

management actions are planned at a stand scale,<br />

and implementation of ecosystem management<br />

approaches will require more attention to habitat<br />

composition and pattern at <strong>the</strong> landscape scale.<br />

Thus, it seems critical to obtain better in<strong>for</strong>mation<br />

at <strong>the</strong>se (as well as o<strong>the</strong>r) scales. The Team’s<br />

ef<strong>for</strong>ts to evaluate habitat use patterns at <strong>the</strong><br />

landscape scale were frustrated by <strong>the</strong> lack of<br />

suitable GIS coverages. Development of such<br />

coverages would greatly facilitate future analyses.<br />

Volume II/Chapter 4<br />

37<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Fur<strong>the</strong>r, any credible attempts to monitor<br />

amounts of spotted owl habitat or trends in such<br />

habitat will require both better data on what<br />

constitutes spotted owl habitat at various spatial<br />

scales, and better data on <strong>for</strong>est structure across<br />

<strong>the</strong> landscape.<br />

Most studies of habitat use-patterns have<br />

been correlative ra<strong>the</strong>r than experimental, which<br />

limits our ability to draw conclusions about<br />

habitat selection by <strong>Mexican</strong> spotted owls.<br />

Fur<strong>the</strong>r, even where correlates of owl occupancy<br />

have been identified, <strong>the</strong>se characteristics have<br />

not been linked to owl fitness. Demographic<br />

studies of <strong>Mexican</strong> spotted owls are underway in<br />

a few areas. Far greater ef<strong>for</strong>ts will be required to<br />

obtain <strong>the</strong> data necessary to determine which<br />

habitats or areas contain self-supporting populations,<br />

and to evaluate habitat composition<br />

within those areas.<br />

In <strong>the</strong> meantime, <strong>for</strong>est management is<br />

ongoing, and we assume that this will continue.<br />

Although controlled experiments in <strong>for</strong>est<br />

management are exceedingly difficult to design<br />

and conduct, <strong>for</strong>est management activities could<br />

provide a great opportunity to learn more about<br />

<strong>the</strong> response of spotted owls to habitat configurations<br />

at various scales. Experimental silvicultural<br />

prescriptions could be developed and<br />

applied, guided by current knowledge of habitat<br />

conditions. Such knowledge should be supplemented<br />

by studies of owl habitat in o<strong>the</strong>r areas<br />

and habitats, and at o<strong>the</strong>r scales. McKelvey and<br />

Wea<strong>the</strong>rspoon (1992) provide an example of a<br />

conceptual approach to integrating silviculture<br />

with knowledge of stand structures used by<br />

spotted owls.<br />

Finally, little attention has been paid to <strong>the</strong><br />

ecology and habitat relationships of <strong>the</strong> owl's<br />

principal prey species (but see Ward and Block<br />

1995), and to how <strong>for</strong>est management might<br />

influence population levels of <strong>the</strong>se species.<br />

Management actions could indirectly affect<br />

spotted owls, ei<strong>the</strong>r positively or negatively,<br />

through effects on <strong>the</strong>ir prey species. There<strong>for</strong>e,<br />

<strong>the</strong>se species should also be considered in future<br />

management planning and research activities.


Volume II/Chapter 4<br />

LITERATURE LITERATURE CITED<br />

CITED<br />

Armstrong, J. A., P. C. Christgau, and K.<br />

Fawcett. 1994. Habitat characteristics of<br />

<strong>Mexican</strong> spotted owl nest sites on <strong>the</strong> Alpine<br />

Ranger District, Apache National Forest,<br />

Arizona. Draft Report. Arizona Game and<br />

Fish Dept., Phoenix. 102pp.<br />

Barrows, C. W. 1981. Roost selection by spotted<br />

owls: an adaptation to heat stress. Condor<br />

83:302-309.<br />

Bendire, C. E. 1892. Life histories of North<br />

American birds. U.S. Nat. Mus. Spec. Bull. 1.<br />

Block, W. M., and L. A. Brennan. 1993. The<br />

habitat concept in ornithology: <strong>the</strong>ory and<br />

application. Current Ornith. 11:35-91.<br />

Brown, D. E., C. H. Lowe, and C. P. Pase. 1980.<br />

A digitized systematic classification <strong>for</strong> ecosystems<br />

with an illustrated summary of <strong>the</strong><br />

natural vegetation of North America. U.S.<br />

For. Serv. Gen. Tech. Rep. RM-73. Rocky<br />

Mtn. For. and Range Exper. Stn. Fort Collins,<br />

Colo. 93pp.<br />

Byers, C. R., R. K. Steinhorst, and P. R.<br />

Krausmann. 1984. Clarification of a technique<br />

<strong>for</strong> analysis of utilization- availability<br />

data. J. Wildl. Manage. 48:1050-1053.<br />

Carey, A. B. 1985. A summary of <strong>the</strong> scientific<br />

basis <strong>for</strong> spotted owl management. Pages 100-<br />

114 in: R. J. Gutiérrez and A. B. Carey, eds.<br />

Ecology and management of <strong>the</strong> spotted owl<br />

in <strong>the</strong> Pacific Northwest. U.S. For. Serv. Gen.<br />

Tech. Rep. PNW-185. Pacific Northwest Res.<br />

Stn. Portland, Oreg.<br />

——., S. P. Horton, and B. L. Biswell. 1992.<br />

Nor<strong>the</strong>rn spotted owls: influence of prey base<br />

and landscape character. Ecol. Monogr.<br />

62:223-250.<br />

Choate, G. A. 1966. New Mexico’s <strong>for</strong>est resource.<br />

U.S. For. Serv. Resour. Bull. INT-5.<br />

Intermountain Res. Stn. Ogden, Utah.<br />

Conner, R. C., J. D. Born, A. W. Green, and R.<br />

A. O’Brien. 1990. Forest resources of Arizona.<br />

U.S. For. Serv. Resour. Bull. INT-69. Intermountain<br />

Res. Stn. Ogden, Utah.<br />

Covington, W. W., and M. M. Moore. 1992.<br />

Postsettlement changes in natural fire regimes:<br />

implications <strong>for</strong> restoration of old-growth<br />

ponderosa pine <strong>for</strong>ests. Pages 81-99 in: M. R.<br />

38<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Kaufmann, W. H. Moir, and R. L. Bassett,<br />

eds. Old-growth <strong>for</strong>ests in <strong>the</strong> Southwest and<br />

Rocky Mountain regions: proceedings of a<br />

workshop. U.S. For. Serv. Gen. Tech. Rep.<br />

RM-213. Rocky Mtn. For. and Range Exper.<br />

Stn. Fort Collins, Colo.<br />

——., and M. M. Moore. 1994. Southwestern<br />

ponderosa <strong>for</strong>est structure: Changes since<br />

Euro-American settlement. J. For. 92:39-47.<br />

Cooper, C. F. 1960. Changes in vegetation,<br />

structure, and growth of southwestern pine<br />

<strong>for</strong>ests since white settlement. Ecol. Monogr.<br />

30:129-164.<br />

Dames and Moore, Inc. 1990. <strong>Spotted</strong> owl<br />

study. Final report submitted to Southwest<br />

Pine Assoc. Dames and Moore, Inc., Tucson,<br />

Ariz. 79pp.<br />

Daniel, T. W., J. A. Helms, and F. S. Baker.<br />

1979. Principles of silviculture. Second ed.<br />

McGraw-Hill, New York, N.Y. 500pp.<br />

Dick-Peddie, W. A. 1993. New Mexico vegetation,<br />

past, present and future. Univ. New<br />

Mexico Press, Albuquerque, N.M. 244pp.<br />

Duncan, R. B., and J. D. Taiz. 1992. A preliminary<br />

understanding of <strong>Mexican</strong> spotted owl<br />

habitat and distribution in <strong>the</strong> Chiricahua<br />

Mountains and associated sub-Mogollon<br />

mountain ranges in sou<strong>the</strong>astern Arizona.<br />

Pages 58-61 in: A. M. Barton and S. A.<br />

Sloane, eds. Proc. Chiricahua Mtns. Res.<br />

Symp. Southwest Parks and Monuments<br />

Assoc., Tucson, Ariz.<br />

Fleischner, T. L. 1994. Ecological costs of<br />

livestock grazing in North America. Conservation<br />

Biology 8:629-644.<br />

Fletcher, K. W. 1990. Habitats used, abundance,<br />

and distribution of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

(<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>) on National Forest<br />

System lands in <strong>the</strong> Southwestern Region.<br />

U.S. For. Serv., Southwestern Region. Albuquerque,<br />

N.M. 56pp.<br />

——., and H. E. Hollis. 1994. Habitats used,<br />

abundance, and distribution of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl (<strong>Strix</strong> <strong>occidentalis</strong> <strong>lucida</strong>) on<br />

National Forest System lands in <strong>the</strong> Southwestern<br />

Region. U.S. For. Serv., Southwestern<br />

Region. Albuquerque, N.M. 86pp.<br />

Ganey, J. L. 1988. Distribution and habitat<br />

ecology of <strong>Mexican</strong> spotted owls in Arizona.


M.S. Thesis, Nor<strong>the</strong>rn Arizona Univ. Flagstaff,<br />

Ariz. 229pp.<br />

——., and R. P. Balda. 1989a. Distribution and<br />

habitat use of <strong>Mexican</strong> spotted owls in Arizona.<br />

Condor 91:355-361.<br />

——., and R. P. Balda. 1989b. Home range<br />

characteristics of spotted owls in nor<strong>the</strong>rn<br />

Arizona. J. Wildl. Manage. 53:1159-1165.<br />

——., and R. P. Balda. 1994. Habitat selection<br />

by <strong>Mexican</strong> spotted owls in nor<strong>the</strong>rn Arizona.<br />

Auk 111:162-169.<br />

——., J. A. Johnson, R. P. Balda, and R. W.<br />

Skaggs. 1988. Status report: <strong>Mexican</strong> spotted<br />

owl. Pages 145-150 in R. L. Glinski et al., eds.<br />

Proc. Southwest Raptor Manage. Symp. and<br />

Workshop. Natl. Wildl. Feder. Sci. and Tech.<br />

Ser. 11.<br />

——., R. B. Duncan, and W. M. Block. 1992.<br />

Use of oak and associated woodlands by<br />

<strong>Mexican</strong> spotted owls in Arizona. Pages 125-<br />

128 in: P. F. Ffolliott, G. J. Gottfried, D. A.<br />

Bennett, V. M. Hernandez, C., A. Ortega-<br />

Rubio, and R. H. Hamre, eds. Ecology and<br />

management of oak and associated woodlands:<br />

perspectives in <strong>the</strong> southwestern United<br />

States and nor<strong>the</strong>rn Mexico. U.S. For. Serv.<br />

Gen. Tech. Rep. RM-218. Rocky Mtn. For.<br />

and Range Exper. Stn. Fort Collins, Colo.<br />

——., R. P. Balda, and R. M. King. 1993.<br />

Metabolic rate and evaporative water loss of<br />

<strong>Mexican</strong> spotted and great horned owls.<br />

Wilson Bull. 105:645-656.<br />

Glover, V. J. 1984. Logging railroads of <strong>the</strong><br />

Lincoln National Forest, New Mexico. U.S.<br />

For. Serv., Southwestern Region. Albuquerque,<br />

N.M.<br />

Gutiérrez, R. J. 1985. An overview of recent<br />

research on <strong>the</strong> spotted owl. Pp. 39-49 in: R.<br />

J. Gutiérrez and A. B. Carey, eds. Ecology and<br />

management of <strong>the</strong> spotted owl in <strong>the</strong> Pacific<br />

Northwest. U.S. For. Serv. Gen. Tech. Rep.<br />

PNW-185. Pacific Northwest Res. Stn.<br />

Portland, Oreg.<br />

——., J. Verner, K. S. McKelvey, B. R. Noon,<br />

G. N. Steger, D. R. Call, W. S. LaHaye, B. B.<br />

Bingham, and J. S. Senser. 1992. Habitat<br />

relations of <strong>the</strong> Cali<strong>for</strong>nia spotted owl. Pages<br />

79-98 in: J. Verner, K. S. McKelvey, B. R.<br />

Noon, R. J. Gutiérrez, G. I. Gould, Jr., and T.<br />

W. Beck, eds. The Cali<strong>for</strong>nia spotted owl: a<br />

Volume II/Chapter 4<br />

39<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

technical assessment of its current status. U.S.<br />

For. Serv. Gen. Tech. Rep. PSW-133. Pacific<br />

Southwest Res. Stn. Albany, Calif.<br />

Harrington, M. G., and S. S. Sackett. 1992. Past<br />

and present fire influences on southwestern<br />

ponderosa pine old-growth. Pp. 44-50 in:<br />

Kauffmann, M. R., W. H. Moir, and R. L.<br />

Bassett, tech. coordinators. Old-growth <strong>for</strong>ests<br />

in <strong>the</strong> southwest and Rocky Mountain regions.<br />

U. S. For. Serv. Gen. Tech. Rep. RM-<br />

213. Rocky Mtn. For. and Range Exper. Stn.<br />

Fort Collins, Colo.<br />

Hull, T. 1995. Comments submitted to <strong>the</strong><br />

Regional Director, U.S. Fish and Wildl. Serv.<br />

regarding <strong>the</strong> draft recovery plan <strong>for</strong> <strong>the</strong><br />

<strong>Mexican</strong> spotted owl, by Applied Ecosystem<br />

Management, Inc. Flagstaff, Ariz.<br />

Hurlbert, S. H. 1984. Pseudoreplication and <strong>the</strong><br />

design of ecological field experiments. Ecol.<br />

Monogr. 54:187-211.<br />

Johnson, J. A., and T. H. Johnson. 1985. The<br />

status of <strong>the</strong> spotted owl in nor<strong>the</strong>rn New<br />

Mexico. New Mexico Dep. Game and Fish,<br />

Santa Fe. 39pp.<br />

——., and T. H. Johnson. 1988. Timber type<br />

model of spotted owl habitat in nor<strong>the</strong>rn New<br />

Mexico. New Mexico Dep. Game and Fish,<br />

Santa Fe. 23pp.<br />

——. 1989. Jemez Mountains spotted owl<br />

survey and habitat evaluation. Santa Fe Natl.<br />

For. Santa Fe, N.M. 19pp.<br />

Johnson, T. H. 1990. Revised timber type model<br />

of spotted owl habitat in nor<strong>the</strong>rn New<br />

Mexico. Santa Fe National Forest. Santa Fe,<br />

N.M. 10pp.<br />

Johnson, M. A. 1995. Changes in Southwestern<br />

<strong>for</strong>ests: stewardship implications. U.S. For.<br />

Serv., Southwestern Region. Albuquerque,<br />

N.M. 6pp.<br />

Kauffman, J. B., and W. C. Krueger. 1984.<br />

Livestock impacts on riparian ecosystems and<br />

streamside management implications, a<br />

review. J. Range Manage. 37:430-438.<br />

Keitt, T. H., A. B. Franklin, and D. L. Urban.<br />

1995. Landscape analysis and metapopulation<br />

structure. Chapter 3 (16 pp.) in: USDI 1995.<br />

<strong>Mexican</strong> spotted owl recovery plan. U. S. Fish<br />

and Wildl. Serv. Albuquerque, NM.<br />

Kennedy, C. E. 1977. Wildlife conflicts in<br />

riparian management: water. Pages 52-58 in:


R. R. Johnson and D. A. Jones, eds. Importance,<br />

preservation and management of<br />

riparian habitat: a symposium. U.S. For. Serv.<br />

Gen. Tech. Rep. RM-43. Rocky Mtn. For.<br />

and Range Exper. Stn. Fort Collins, Colo.<br />

Kertell, K. 1977. The spotted owl at Zion<br />

National Park, Utah. West. Birds 8:147-150.<br />

Knopf, F. L., R. R. Johnson, T. Rich, F. B.<br />

Samson, and R. C. Szaro. 1988. Conservation<br />

of riparian ecosystems in <strong>the</strong> United States.<br />

Wilson Bull. 100:272-284.<br />

Kroel, K. W. 1991. Home range and habitat use<br />

characteristics of <strong>the</strong> <strong>Mexican</strong> spotted owls<br />

(sic) in <strong>the</strong> sou<strong>the</strong>rn Sacramento Mountains,<br />

New Mexico. M.S. Thesis. New Mexico State<br />

Univ., Las Cruces. 63pp.<br />

Ligon, J. S. 1926. Habits of <strong>the</strong> spotted owl<br />

(Syrnium occidentale). Auk 43:421-429.<br />

Madany, M. H., and N. E. West. 1983. Livestock<br />

grazing-fire regime interactions within<br />

montane <strong>for</strong>ests of Zion National Park, Utah.<br />

Ecology 64:661-667.<br />

McDonald, C. B., J. Anderson, J. C. Lewis, R.<br />

Mesta, A. Ratzlaff, T. J. Tibbitts, and S. O.<br />

Williams. 1991. <strong>Mexican</strong> spotted owl (<strong>Strix</strong><br />

<strong>occidentalis</strong> <strong>lucida</strong>) status report. U.S. Fish and<br />

Wildl. Serv. Albuquerque, N.M. 85pp.<br />

McKelvey, K. S., and C. P. Wea<strong>the</strong>rspoon. 1992.<br />

Projected trends in owl habitat. Pages 261-<br />

275 in: J. Verner, K. S. McKelvey, B. R.<br />

Noon, R. J. Gutiérrez, G. I. Gould, Jr., and T.<br />

W. Beck, eds. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status. U.S.<br />

For. Serv. Gen. Tech. Rep. PSW-133. Pacific<br />

Southwest Res. Stn. Albany, Calif.<br />

Minckley, W. L., and T. O. Clark. 1984. Formation<br />

and destruction of a Gila River mesquite<br />

bosque community. Desert <strong>Plan</strong>ts 6:23-30.<br />

——., and J. N. Rinne. 1985. Large woody<br />

debris in hot-desert streams: an historical<br />

review. Desert <strong>Plan</strong>ts 7:142- 153.<br />

Mohr, C. O. 1947. Table of equivalent populations<br />

of North American small mammals.<br />

Amer. Midl. Nat. 37:223-249.<br />

Neu, C. W., C. R. Byers, and J. M. Peek. 1974.<br />

A technique <strong>for</strong> analysis of utilization-availability<br />

data. J. Wildl. Manage. 38:541-545.<br />

O’Neill, R. V., B. T. Milne, M. G. Turner, and<br />

R. H. Gardner. 1988. Resource utilization<br />

Volume II/Chapter 4<br />

40<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

scales and landscape pattern. Landscape Ecol.<br />

2:63-69.<br />

Orians, G. H., and J. F. Wittenberger. 1991.<br />

Spatial and temporal scales in habitat selection.<br />

Am. Nat. 137(Supplement):s29-s49.<br />

Phillips, A. R., J. T. Marshall, Jr., and G.<br />

Monson. 1964. The birds of Arizona. Univ. of<br />

Arizona Press, Tucson. 212pp.<br />

Platts, W. S. 1990. Managing fisheries and<br />

wildlife on rangelands grazed by livestock: a<br />

guidance and reference document <strong>for</strong> biologists.<br />

Nevada Dep. of Wildl. 448pp.<br />

Porter, W. F., and K. E. Church. 1987. Effects of<br />

environmental pattern on habitat preference<br />

analysis. J. Wildl. Manage. 51:681-685.<br />

Reynolds, R. T. 1993. Ecology and dispersal of<br />

<strong>the</strong> <strong>Mexican</strong> spotted owl in Colorado, FY<br />

1993-94. Research Proposal. U.S. For. Serv.<br />

Rocky Mtn. For. and Range Exper. Stn. Fort<br />

Collins, Colo.<br />

Rinkevich, S. E. 1991. Distribution and habitat<br />

characteristics of <strong>Mexican</strong> spotted owls in<br />

Zion National Park, Utah. M.S. Thesis,<br />

Humboldt State Univ., Arcata, Calif. 62pp.<br />

——., and R. J. Gutiérrez. In review. <strong>Mexican</strong><br />

spotted owl habitat characteristics in Zion<br />

National Park. J. Raptor Res.<br />

Ruess, B. J. 1995. Changes in <strong>Mexican</strong> spotted<br />

owl habitat within ponderosa pine/Gambel<br />

oak communities since Euro-American<br />

settlement. M.S. Thesis. Nor<strong>the</strong>rn Arizona<br />

University, Flagstaff. 42 pp.<br />

Rummel, R. S. 1951. Some effects of livestock<br />

grazing on ponderosa pine <strong>for</strong>est and range in<br />

central Washington. Ecology 32:594-607.<br />

Savage, M., and T. W. Swetnam. 1990. Early<br />

19th century fire decline following sheep<br />

pasturing in a Navajo ponderosa pine <strong>for</strong>est.<br />

Ecology 32:594-607.<br />

Schulz, T. T., and W. C. Leininger. 1990. Differences<br />

in riparian vegetation structure between<br />

grazed areas and exclosures. J. Range Manage.<br />

43:295-299.<br />

——., and W. C. Leininger. 1991. Nongame<br />

wildlife communities in grazed and ungrazed<br />

montane riparian sites. Great Basin Naturalist<br />

51:286-292.<br />

Seamans, M. E., and R. J. Gutiérrez. In press.<br />

Breeding habitat ecology of <strong>the</strong> <strong>Mexican</strong>


spotted owl in <strong>the</strong> Tularosa Mountains, New<br />

Mexico. Condor.<br />

Skaggs, R. W. 1988. Status of <strong>the</strong> spotted owl in<br />

sou<strong>the</strong>rn New Mexico: 1900-1987. New<br />

Mexico Dep. Game and Fish, Santa Fe. 13pp.<br />

——., and R. J. Raitt. 1988. A spotted owl<br />

inventory of <strong>the</strong> Lincoln National Forest,<br />

Sacramento Division, 1988. New Mexico<br />

Dep. Game and Fish, Santa Fe. 12pp.<br />

Skovlin, J. M. 1984. Grazing impacts on wetlands<br />

and riparian habitat: a review of our<br />

knowledge. Pages 1001-1004 in National<br />

Research Council and National Academy of<br />

Sciences. Developing strategies <strong>for</strong> rangeland<br />

management: a report prepared by <strong>the</strong> committee<br />

on developing strategies <strong>for</strong> rangeland<br />

management. Westview Press, Boulder, Colo.<br />

Spencer, J. S., Jr. 1966. Arizona’s <strong>for</strong>ests. U. S.<br />

For. Serv. Resourc. Bull. INT-6. Intermountain<br />

Res. Stn. Ogden, Utah.<br />

Stein, S. J. 1988. Explanations of <strong>the</strong><br />

imbalanced age structure and scattered distribution<br />

of ponderosa pine within a highelevation<br />

mixed coniferous <strong>for</strong>est. For. Ecol.<br />

and Manage. 25:139-153.<br />

SWCA, Inc. 1992. Nest tree and nest tree<br />

habitat characteristics of <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl in Arizona and New Mexico. Final Rep.<br />

submitted to U.S. For. Serv., Southwestern<br />

Region. Albuquerque, N.M. 31pp.<br />

Tarango, L. A., R. Valdez, and P. J. Zwank.<br />

1994. <strong>Mexican</strong> spotted owl distribution and<br />

habitat characterizations in southwestern<br />

Chihuahua, Mexico. Final report, Contract<br />

No. 80-516.6-66. New Mexico Dep. Game<br />

and Fish, Santa Fe. 69pp.<br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner. 1990. A<br />

conservation strategy <strong>for</strong> <strong>the</strong> spotted owl. U.S.<br />

Gov. Print. Off., Washington, D.C. 427pp.<br />

Turner, M. G. 1989. Landscape ecology: <strong>the</strong><br />

effect of pattern on process. Annu. Rev. Ecol.<br />

Syst. 20:171-197.<br />

Urban, D. L., R. V. O’Neill, and H. H. Shugart,<br />

Jr. 1987. Landscape ecology: a hierarchical<br />

perspective can help scientists understand<br />

spatial patterns. BioScience 37:119-127.<br />

USDA Forest Service. 1993. Facts about <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. U. S. For. Serv. Southwestern<br />

Region, Albuquerque, N.M.<br />

Volume II/Chapter 4<br />

41<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

USDI Fish and Wildlife Service. 1993. Endangered<br />

and threatened wildlife and plants; final<br />

rule to list <strong>the</strong> <strong>Mexican</strong> spotted owl as a<br />

threatened species. Federal Register 58:<br />

14248-14271.<br />

——. 1995. <strong>Mexican</strong> spotted owl <strong>Recovery</strong><br />

<strong>Plan</strong>. Albuquerque, N.M.<br />

USGS. 1904. Forest conditions in <strong>the</strong> San<br />

Francisco Mountains Reserve. U. S. Geological<br />

Survey Prof. Paper 22. Washington, D.C.<br />

Utah <strong>Mexican</strong> spotted owl technical team. 1994.<br />

Suggestions <strong>for</strong> management of <strong>Mexican</strong><br />

spotted owls in Utah. Utah Div. Wildl.<br />

Resources. Salt Lake City, UT. 103pp.<br />

Van Hooser, D. D., R. A. O’Brien, and D. C.<br />

Collins. 1993. New Mexico’s <strong>for</strong>est resources.<br />

U.S. For. Serv. Resour. Bull. INT-79.<br />

Intermtn. Res. Stn., Ogden, Utah. 80pp.<br />

Verner, J., R. J. Gutiérrez, and G. I. Gould, Jr.<br />

1992. The Cali<strong>for</strong>nia spotted owl: general<br />

biology and ecological relations. Pp. 55-78 in:<br />

Verner, J., K. S. McKelvey, B. R. Noon, R. J.<br />

Gutiérrez, G. I. Gould, Jr., and T. W. Beck,<br />

eds. The Cali<strong>for</strong>nia spotted owl: a technical<br />

assessment of its current status. U.S. For. Serv.<br />

Gen. Tech. Rep. PSW-133. Pacific Southwest<br />

Res. Stn. Albany, Calif.<br />

Ward, J. P., Jr., and W. M. Block. 1995. <strong>Mexican</strong><br />

spotted owl prey ecology. Chapter 5 (48 pp.)<br />

in: USDI Fish and Wildlife Service. 1995.<br />

<strong>Mexican</strong> spotted owl recovery plan.<br />

Albuquerque, N.M.<br />

Willey, D. W. 1992. Movements and habitat<br />

ecology of <strong>Mexican</strong> spotted owls in sou<strong>the</strong>rn<br />

Utah. Final report. Utah Div. Wildl. Resources,<br />

Salt Lake City. 23pp.<br />

——. 1993. Home-range characteristics and<br />

juvenile dispersal ecology of <strong>Mexican</strong> spotted<br />

<strong>Owl</strong>s in sou<strong>the</strong>rn Utah. Rep. submitted to<br />

Utah Div. Wildl. Resources, Salt Lake City.<br />

44pp.<br />

Woodhouse, S. W. 1853. Report on <strong>the</strong> natural<br />

history of <strong>the</strong> country passed over by <strong>the</strong><br />

exploring expedition under <strong>the</strong> command of<br />

Brev. Capt. L. Sitgreaves, U.S. Topographical<br />

Engineers, during <strong>the</strong> year 1851. [and]<br />

Zoology: Birds. Pages 33-40, 58-105 in Capt.<br />

L. Sitgreaves, ed. Report of an expedition<br />

down <strong>the</strong> Zuni and Colorado Rivers. Robert<br />

Armstrong, Public Printer, Washington, D.C.


Woolsey, T. S., Jr. 1911. Western yellow pine in<br />

Arizona and New Mexico. U. S. For. Serv.<br />

Bull. 101. Washington, D.C.<br />

Worton, B. J. 1989. Kernel methods <strong>for</strong> estimating<br />

<strong>the</strong> utilization distribution in home-range<br />

studies. Ecology 70:164-168.<br />

Zabel, C., K. McKelvey, and J. P. Ward, Jr. 1995.<br />

Influence of primary prey on home-range size<br />

and habitat-use patterns of <strong>Spotted</strong> <strong>Owl</strong>s<br />

(<strong>Strix</strong> <strong>occidentalis</strong>). Can. J. Zool. 73:433-439.<br />

Zhou, B. 1994. Quantifying <strong>Mexican</strong> spotted<br />

owl habitat with <strong>the</strong> Johnson system of<br />

distributions. M.S. Thesis. Nor<strong>the</strong>rn Arizona<br />

Univ., Flagstaff. 169pp.<br />

Zimmerman, G. T., and L. F. Neuenschwander.<br />

1984. Livestock grazing influences on community<br />

structure, fire intensity, and fire<br />

frequency within <strong>the</strong> Douglas-fir/ninebark<br />

habitat type. J. Range Manage. 37:104-110.<br />

Zwank, P. J., K. W. Kroel, D. M. Levin, G. M.<br />

Southward, and R. C. Rommé. 1994. Habitat<br />

characteristics of <strong>Mexican</strong> spotted owls in<br />

southwestern New Mexico. J. Field Ornithol.<br />

65: 324-334.<br />

Volume II/Chapter 4<br />

42<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


Volume II/Chapter 5<br />

CHAPTER CHAPTER 5: 5: <strong>Mexican</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Owl</strong> Prey Prey Ecology<br />

Ecology<br />

In addition to shelter, water, and o<strong>the</strong>r<br />

requirements, habitat must also provide spotted<br />

owls with food. Certain trees within specific<br />

<strong>for</strong>est communities may meet nesting, roosting,<br />

and perching needs, but <strong>the</strong> tree component<br />

alone may not necessarily sustain <strong>the</strong> animal<br />

species upon which <strong>the</strong> owls prey. Conserving<br />

appropriate habitat <strong>for</strong> <strong>the</strong> owl includes conserving<br />

habitat <strong>for</strong> a suite of prey species.<br />

The distribution and abundance of prey<br />

often influence <strong>the</strong> distribution, abundance, and<br />

reproduction of raptors (Newton 1979). In<br />

owls, reproductive success is often correlated<br />

with prey abundance (Craighead and Craighead<br />

1956, Sou<strong>the</strong>rn 1970, Lundberg 1976,<br />

Wendland 1984, Korpimäki and Norrdahl<br />

1991). The postulated mechanism behind this<br />

relationship is energetically based. Male owls<br />

must provide enough food to <strong>the</strong>ir female mates<br />

during incubation and brooding to prevent<br />

abandonment of nests or young (Johnsgard<br />

1988). Accordingly, ecologists suspect that<br />

spotted owls select habitats partially because of<br />

<strong>the</strong> availability of prey (Carey 1985, Thomas et<br />

al. 1990, Verner et al. 1992). Understanding <strong>the</strong><br />

natural history of <strong>the</strong> spotted owl’s primary prey<br />

is vital in<strong>for</strong>mation <strong>for</strong> <strong>the</strong> <strong>Recovery</strong> <strong>Plan</strong> because<br />

it provides resource planners and managers<br />

with ano<strong>the</strong>r tool <strong>for</strong> evaluating an area’s ability<br />

to support spotted owls.<br />

This section summarizes in<strong>for</strong>mation about<br />

spotted owl-prey relationships and <strong>the</strong> ecology of<br />

<strong>the</strong> owl’s prey. Specifically, our objectives are to:<br />

(1) describe <strong>the</strong> diet of <strong>the</strong> <strong>Mexican</strong> spotted owl;<br />

(2) identify prey that may influence owl fitness;<br />

and (3) quantify habitat correlates of <strong>the</strong> owl’s<br />

primary prey.<br />

Ideally, relationships among <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, its prey, and <strong>the</strong> prey’s habitat<br />

should be examined across different spatial and<br />

temporal scales. In reality, <strong>the</strong> available in<strong>for</strong>mation<br />

permits only a limited view of <strong>the</strong> owl’s prey<br />

ecology. For example, we could describe abundance<br />

and distribution of <strong>the</strong> owl’s common<br />

prey among different vegetation communities,<br />

but could not provide a direct link between owl<br />

James P. Ward, Jr., and William M. Block<br />

1<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

habitat use and prey availability. Although <strong>the</strong><br />

latter in<strong>for</strong>mation is preferred <strong>for</strong> prioritizing<br />

habitat conservation, our approach relies upon<br />

conserving a general mixture of habitats <strong>for</strong> <strong>the</strong><br />

owl and its prey throughout major portions of<br />

<strong>the</strong> owl’s range. In time, in<strong>for</strong>mation from more<br />

specific studies should be used to refine <strong>the</strong><br />

general findings presented here.<br />

METHODS METHODS FOR FOR<br />

FOR<br />

DETERMINING DETERMINING OWL OWL DIETS<br />

DIETS<br />

The dietary habits of raptors can be determined<br />

both directly and indirectly. Observations<br />

of prey capture and of prey taken to roosts or<br />

nests provide direct evidence of a raptor’s diet.<br />

However, such observations are difficult to<br />

obtain from nocturnal <strong>for</strong>agers like owls and<br />

offer little opportunity <strong>for</strong> quantitative analysis.<br />

Regurgitated pellets of undigested materials (fur,<br />

fea<strong>the</strong>rs, bones, chitinous exoskeletons) offer an<br />

indirect, alternative method <strong>for</strong> analyzing <strong>the</strong><br />

feeding habits of owls (Errington 1930, Glading<br />

et al. 1943, Marti 1987).<br />

<strong>Spotted</strong> owl prey are identified by examining<br />

<strong>the</strong> contents of pellets collected below roosts and<br />

nests. Prey remains are identified to species,<br />

genus, or a less specific prey category and tallied.<br />

Diets are <strong>the</strong>n quantified using two measures,<br />

relative frequency and percent biomass (Forsman<br />

et al. 1984, Marti 1987). Relative frequency of<br />

prey is expressed as <strong>the</strong> number of individual<br />

items of a given prey species or group divided by<br />

<strong>the</strong> total number of all individual items found in<br />

a sample. Biomass (g) is <strong>the</strong> total number of<br />

items of a given prey species or group multiplied<br />

by <strong>the</strong> average mass (g) <strong>for</strong> that species or group.<br />

Commonly, both relative frequency and biomass<br />

are expressed as percentages. Both measures can<br />

be used to compare owl diets among sampling<br />

units such as reproductive groups, locations,<br />

seasons, and so on. The two measures provide<br />

different in<strong>for</strong>mation. Measures of relative<br />

frequency indicate <strong>the</strong> proportion of each prey<br />

type in <strong>the</strong> owl’s diet by number, whereas,


percent biomass indicates <strong>the</strong> proportion of each<br />

prey type by weight.<br />

Pellets may provide biased measures of owl<br />

diets. Pellets are typically collected opportunistically<br />

below roosting owls or from known roost<br />

groves, not randomly or systematically. The bias<br />

potentially resulting from nonrandom sampling<br />

could not be evaluated. We assumed that items<br />

found in pellets reflected <strong>the</strong> true proportions of<br />

prey species in owl diets. The methods we used<br />

to quantify <strong>Mexican</strong> spotted owl diet are comparable<br />

to o<strong>the</strong>r studies conducted on <strong>the</strong> nor<strong>the</strong>rn<br />

and Cali<strong>for</strong>nia subspecies (see reviews by Thomas<br />

et al. 1990 and Verner et al. 1992). Accordingly,<br />

we restricted analyses and inferences to<br />

relative comparisons.<br />

We compiled and analyzed prey remains of<br />

<strong>Mexican</strong> spotted owls as reported in 13 studies<br />

conducted since 1977 (Tables 5.1-5.6). A total<br />

of 11,164 prey items was examined. These data<br />

consisted of 25 data sets from 18 geographic<br />

areas throughout <strong>the</strong> owl’s range, and included<br />

most published and unpublished in<strong>for</strong>mation of<br />

<strong>Mexican</strong> spotted owl diet through 1993 (Tables<br />

5.1-5.6). Kertell (1977) was not used because of<br />

<strong>the</strong> small number of items reported.<br />

Each data set differed in <strong>the</strong> number of owls,<br />

years, and pellets examined. In some cases, <strong>the</strong><br />

number of owls studied or pellets collected were<br />

not recorded. Thus, we could not use owls as <strong>the</strong><br />

sampling unit or separate differences among owl<br />

territories in region-wide analyses. However,<br />

each data set contained a known number of<br />

identified prey items. For analysis, we treated<br />

each prey item as an observation and each data<br />

set as a dietary sample. Following this logic, <strong>the</strong><br />

number of observations in a sample (i.e., sample<br />

size) corresponds to <strong>the</strong> total number of prey<br />

items identified in each data set. We justified <strong>the</strong><br />

use of prey items as an observational unit instead<br />

of pellets because a pellet is difficult to define<br />

(i.e., broken pellets are often collected or multiple<br />

pellets are stored toge<strong>the</strong>r and break apart<br />

during storage) and a single pellet may contain<br />

multiple prey items. Being an uncertain and<br />

inconsistent measure of sample size, no pellet<br />

total is given here.<br />

The methods used to identify remains and<br />

tally prey numbers followed Forsman et al.<br />

(1984). Samples abbreviated as CAPRF2,<br />

Volume II/Chapter 5<br />

2<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

ZION2, CANYL, BAR-M, MOGAZ2,<br />

COCODM, GILADM, and SACMT2 (see<br />

Tables 5.1-5.6 <strong>for</strong> acronym definition) were<br />

analyzed using standardized procedures described<br />

and implemented by DeRosier and Ward<br />

(1994) to facilitate comparison. According to<br />

both sources, remains were keyed to species<br />

when possible using skulls and appendicular<br />

skeletal parts. Specialists were consulted to<br />

identify less common or unusual remains,<br />

particularly bats, birds, reptiles, and invertebrates.<br />

Deviations from our standard identification<br />

procedures were necessary <strong>for</strong> <strong>the</strong> SCCOL<br />

sample where invertebrate parts were not identified<br />

(Charles Johnson, Rocky Mountain Research<br />

Station, Fort Collins, CO, pers. comm.),<br />

and in <strong>the</strong> ZION1 sample where appendicular<br />

parts were not used to tally prey (Sarah<br />

Rinkevich, FWS, Albuquerque, NM, pers.<br />

comm.).<br />

<strong>Owl</strong> diets were quantified using relative<br />

frequency and percent biomass <strong>for</strong> 11 prey<br />

groups: woodrats, white-footed (peromyscid)<br />

mice, voles, pocket gophers, rabbits, bats, o<strong>the</strong>r<br />

or unidentified small mammals (mostly murids),<br />

o<strong>the</strong>r or unidentified medium mammals (mostly<br />

sciurids), birds, reptiles, and arthropods (Tables<br />

5.1-5.6). We use <strong>the</strong> term peromyscid mice here<br />

in place of white-footed mice to represent <strong>the</strong><br />

approximate 15 North American species of <strong>the</strong><br />

genus Peromyscus. Confusion sometimes follows<br />

discussion of white-footed mice because this is<br />

also <strong>the</strong> common name <strong>for</strong> P. leucopus.<br />

Several sources were used to estimate prey<br />

mass (Appendix 5a). Biomass estimates given by<br />

<strong>the</strong> original authors were retained unless better<br />

estimates were available. Estimates of prey mass<br />

from areas nearest to where pellets were collected<br />

were used whenever possible. In <strong>the</strong> absence of<br />

better data, general references <strong>for</strong> mass were<br />

used. Averages, weighted according to proportions<br />

of species in owl diets, were used to estimate<br />

biomass <strong>for</strong> less specific taxa such as<br />

“woodrat species” or “unidentified bat.”


3<br />

Table Table 5.1. 5.1. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> subspecies' range. Values<br />

were calculated from totals pooled across owl territories and years.


4<br />

Table Table 5.2. 5.2. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> central portion of <strong>the</strong> subspecies' range. Values<br />

were calculated from totals across owl territories and years.


5<br />

Table Table 5.3. 5.3. Relative frequency of prey items found in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> sou<strong>the</strong>rn portion of <strong>the</strong> subspecies' range. Values<br />

were calculated from totals pooled across owl territories and years.


6<br />

Table Table 5.4. 5.4. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> subspecies' range. Values were calculated<br />

from totals pooled across owl territories and years.


7<br />

Table Table 5.5. 5.5. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> central portion of <strong>the</strong> subspecies' range. Values were calculated<br />

from totals pooled across owl territories and years.


8<br />

Table Table 5.6. 5.6. Percent of prey biomass in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls occurring in <strong>the</strong> sou<strong>the</strong>rn portion of <strong>the</strong> subspecies' range. Values were calculated<br />

from totals pooled across owl territories and years.


GENERAL GENERAL GENERAL FOOD FOOD HABITS<br />

HABITS<br />

The <strong>Mexican</strong> spotted owl eats a variety of<br />

animals throughout its range but usually takes<br />

small and medium-sized mammals (Table 5.7).<br />

As a group, mammals are taken more frequently<br />

(x = 82.1%; cv = 13.3%; n = 25 data sets) than<br />

birds (x = 4.8%; cv = 78.4%; n = 25), reptiles<br />

(x = 0.5%; cv = 192.5%; n = 25) or arthropods<br />

(x = 12.6%; cv = 79.0%, n = 23). Arthropods<br />

could be present in <strong>the</strong> owl’s diet because <strong>the</strong>se<br />

items were eaten by <strong>the</strong> owl’s prey prior to<br />

capture. Mammal use is even greater when<br />

biomass is considered (x = 95.8%, x = 3.9% <strong>for</strong><br />

birds, x = 0.1% <strong>for</strong> reptiles, and x = 0.2% <strong>for</strong><br />

arthropods). A cumulative plot of diet frequencies<br />

(Figure 5.1a) indicates that 90% of an<br />

“average” <strong>Mexican</strong> spotted owl diet would<br />

contain 30% woodrats; 28% peromyscid mice;<br />

13% arthropods; 9% microtine voles; 5% birds;<br />

and 4% medium-sized rodents, mostly diurnal<br />

sciurids. A cumulative plot of diet biomass<br />

(Figure 5.1b) indicates that, on average, 90% of<br />

a <strong>Mexican</strong> spotted owl diet is comprised of 53%<br />

woodrats; 13% rabbits; 9% peromyscid mice;<br />

9% birds; and 6% medium-sized mammals such<br />

as diurnal sciurids. These rangewide patterns,<br />

however, are not consistent among RUs.<br />

To evaluate geographic variation of <strong>Mexican</strong><br />

spotted owl food habits, we examined diet data<br />

by recovery unit. Frequencies of prey were<br />

treated as binomial, ei<strong>the</strong>r <strong>the</strong> prey species being<br />

compared was consumed during a successful<br />

<strong>for</strong>aging event or some o<strong>the</strong>r species was consumed.<br />

We trans<strong>for</strong>med <strong>the</strong>se data <strong>for</strong> parametric<br />

analyses using an arcsine-squareroot trans<strong>for</strong>mation<br />

(Freeman and Tukey 1950) and tested<br />

<strong>the</strong> hypo<strong>the</strong>sis of no difference in diets among<br />

RUs using a one-way analysis of variance<br />

(ANOVA). A Tukey’s test was used to conduct<br />

multiple comparisons among RUs when <strong>the</strong><br />

ANOVA was significant. We conducted <strong>the</strong> tests<br />

<strong>for</strong> each of <strong>the</strong> 11 prey groups separately. We did<br />

not conduct an ANOVA on percent biomass<br />

estimates because <strong>the</strong>se data were comprised of<br />

both frequency and prey mass, and our estimates<br />

of mass <strong>for</strong> prey were not significantly different<br />

among RUs except <strong>for</strong> voles (F = 4.4, d.f. = 6,<br />

15, P = 0.009). Thus, prey mass would have<br />

Volume II/Chapter 5<br />

9<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

been roughly equivalent among RUs and comparisons<br />

of percent biomass would have been<br />

equivalent to those of relative frequency.<br />

Our analyses indicated significant differences<br />

in <strong>Mexican</strong> spotted owl diets among geographic<br />

location (Figure 5.2). Woodrats were taken more<br />

often in <strong>the</strong> Colorado Plateau compared to <strong>the</strong><br />

Upper Gila Mountains and Sierra Madre<br />

Occidentalis - Norte (F = 4.6, d.f. = 6, 18,<br />

P = 0.005; Figure 5.2a). Voles were consumed<br />

more frequently by owls in <strong>the</strong> Basin and<br />

Range - East, Sou<strong>the</strong>rn Rocky Mountains -<br />

Colorado, and Upper Gila Mountains RUs than<br />

in <strong>the</strong> Colorado Plateau and Basin and<br />

Range - West RUs (F = 8.3, d.f. = 6, 18,<br />

P < 0.001; Figure 5.2a). Pocket gophers were<br />

eaten more often in <strong>the</strong> Upper Gila Mountains<br />

compared to <strong>the</strong> Colorado Plateau (F = 4.5,<br />

d.f. = 6, 18, P = 0.006; Figure 5.2c); and birds<br />

were consumed more often by owls in <strong>the</strong><br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico,<br />

Basin and Range - West, and Upper Gila Mountains<br />

RUs than in <strong>the</strong> Colorado Plateau (F = 5.8,<br />

d.f. = 6, 18, P = 0.002; Figure 5.2d). Bats and<br />

reptiles were consumed more frequently in <strong>the</strong><br />

Basin and Range - West RU compared to <strong>the</strong><br />

Upper Gila Mountains RU (F = 2.9, d.f. = 6, 18,<br />

P = 0.038; Figure 5.2e and F = 3.3, d.f. = 6, 18,<br />

P = 0.024; Figure 5.2f, respectively). Of <strong>the</strong> five<br />

groups that did not differ significantly among<br />

RUs (peromyscid mice, rabbits, o<strong>the</strong>r small<br />

mammals, o<strong>the</strong>r medium mammals, and<br />

arthropods), only peromyscid mice (x = 27.2%,<br />

cv = 42% among units) and arthropods (x =<br />

13.8%, cv = 70%) were frequent food items.<br />

Interregional trends in prey consumption<br />

could be attributed to o<strong>the</strong>r factors like temporal<br />

variation in <strong>the</strong> owl’s diet within each study area,<br />

or because <strong>the</strong> breeding status of owls differed<br />

among studies. Diet frequencies from each study<br />

were pooled across years and owl pairs regardless<br />

of reproductive status because this in<strong>for</strong>mation<br />

was unavailable in most cases. However, <strong>the</strong> year<br />

in which diet samples were collected and <strong>the</strong><br />

reproductive success of a sufficient number of<br />

owls were known <strong>for</strong> three studies (SACMT2,<br />

GILADM, COCODM). These three data sets<br />

were collected from two of <strong>the</strong> RUs, Basin and<br />

Range - East (SACMT2; Tables 5.3 and 5.6) and<br />

Upper Gila Mountains (GILADM, COCODM;


Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Table Table 5.7. 5.7. Animal species consumed by <strong>Mexican</strong> spotted owls as determined from examination of<br />

25 data sets collected from 18 geographic areas.<br />

FPO<br />

10


Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure Figure 5.1. 5.1. 5.1. Cumulative distributions of prey in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls presented as (a) (a)<br />

(a)<br />

relative frequency (%) of items and (b) (b) percent biomass. Values are averages across 25 data sets conducted<br />

throughout <strong>the</strong> owl's range and dashed lines are 95% confidence limits. Prey groups are<br />

WRAT-woodrats; MICE-peromyscid mice; ARTH-arthropods; VOLE-voles; BIRD-birds; OMMMo<strong>the</strong>r<br />

medium-sized mammals; BATS-bats; OSMM-o<strong>the</strong>r small-sized mammals; RABB-rabbits;<br />

GOPH-gophers; REPT-reptiles.<br />

11


Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure 5.2. 5.2. Geographic variability in <strong>the</strong> food habits of <strong>Mexican</strong> spotted owls presented as relative<br />

frequencies of (a) (a) woodrats, (b) (b) voles, (c) (c) gophers, (d) (d) birds, (e) (e) bats, and (f) (f) reptiles. Point values are<br />

from single studies or averages among <strong>the</strong> number of data sets shown in paren<strong>the</strong>sis. Vertical bars are<br />

95% confidence intervals showing sampling and inter-data set variation within a recovery unit. <strong>Recovery</strong><br />

Unit acronyms are COPLAT-Colorado Plateau; SRM-CO-Sou<strong>the</strong>rn Rocky Mountains - Colorado;<br />

SRM-NM-Sou<strong>the</strong>rn Rocky Mountains - New Mexico; UPGIL-Upper Gila Mountains; BAR-W-<br />

Basin and Range - West; BAR-E-Basin and Range - East; and SMO-N-Sierra Madre Occidental -<br />

Norte.<br />

12


Tables 5.2 and 5.5). <strong>Owl</strong> reproductive success<br />

was determined similarly in all three studies<br />

using methods described by Forsman (1983). We<br />

used data from <strong>the</strong>se studies to evaluate <strong>the</strong><br />

influence of geography, time, and owl reproductive<br />

status on <strong>the</strong> interregional trends in <strong>the</strong> owl’s<br />

diet.<br />

To quantify <strong>the</strong> effect of <strong>the</strong>se three factors,<br />

we analyzed <strong>the</strong> differences in relative frequency<br />

of a given prey group in <strong>the</strong> owl’s diet among<br />

study area (SACMT2, COCODM, or<br />

GILADM; geographic variation), year (1991-<br />

1993; temporal variation), and number of owl<br />

young produced (0, 1, 2, or 3; variation in owl<br />

reproductive status) using a three-factor<br />

ANOVA. Prey groups that were common to<br />

owls in any of <strong>the</strong> three study areas were analyzed<br />

separately (woodrats, peromyscid mice,<br />

voles, pocket gophers, rabbits, o<strong>the</strong>r mediumsized<br />

mammals, and arthropods). We define<br />

common prey arbitrarily as those species contributing<br />

>10% of diet frequency or biomass, which<br />

includes <strong>the</strong> majority of species identified in <strong>the</strong><br />

90% cumulative averages (Figure 5.1). Statistical<br />

significance of each factor was used to quantify<br />

<strong>the</strong> importance of <strong>the</strong>se factors in determining<br />

regional diet trends.<br />

In 3 of 7 tests, consumption of common<br />

prey varied primarily by geographic location<br />

(Table 5.8). In <strong>the</strong> o<strong>the</strong>r 4 tests, consumption of<br />

woodrats, peromyscid mice, medium-sized<br />

mammals, and arthropods was influenced by an<br />

interaction between geographic location and<br />

year. However, significance values indicated that<br />

<strong>the</strong> consumption of woodrats and arthropods<br />

was influenced more by <strong>the</strong> owls’ location than<br />

<strong>the</strong> particular year (Table 5.8). Prevalence of<br />

geographic variation in <strong>the</strong> three-factor ANOVA<br />

results supports our claims that <strong>the</strong> owl’s feeding<br />

habitats differ among regions, as discussed above<br />

(Figure 5.2).<br />

Diet differences of <strong>the</strong> <strong>Mexican</strong> spotted owl<br />

likely result from a combination of habitat<br />

differences among RUs <strong>for</strong> both <strong>the</strong> owl and its<br />

prey (See Zoogeography and Macrohabitats of<br />

Common Prey). Landscape approaches to<br />

management that maintain conditions <strong>for</strong><br />

common prey of any predator are assumed to<br />

have beneficial effects (Reynolds et al. 1992).<br />

However, <strong>the</strong> reliability of such conservation<br />

Volume II/Chapter 5<br />

13<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

strategies must be firmly based on ecological<br />

links among prey, predator, and environmental<br />

conditions. Thus, it is crucial to consider relationships<br />

among prey abundance and persistence<br />

of owl populations, and among prey abundance,<br />

availability, and habitat conditions.<br />

RELATIVE RELATIVE IMPORTANCE<br />

IMPORTANCE<br />

OF OF PREY<br />

PREY<br />

Prey that positively influences owl survival,<br />

reproduction, or numbers may increase <strong>the</strong><br />

likelihood of persistence of <strong>Mexican</strong> spotted owl<br />

populations. Although no in<strong>for</strong>mation is available<br />

to quantify effects of food on spotted owl<br />

survival and density, previous studies have<br />

examined correlates between this owl’s diet and<br />

reproduction. For example, Barrows (1987)<br />

suggested that larger prey (e.g., woodrats) was<br />

taken in greater frequency by owls with young.<br />

Thrailkill and Bias (1989) reported a similar<br />

pattern <strong>for</strong> Cali<strong>for</strong>nia spotted owls occurring in<br />

<strong>the</strong> central Sierra Nevada. In contrast, Ward<br />

(1990) observed a different pattern <strong>for</strong> nor<strong>the</strong>rn<br />

spotted owls in northwestern Cali<strong>for</strong>nia. He<br />

found that large prey was taken in relatively<br />

equal frequency by breeding and nonbreeding<br />

owls, presumably because woodrats were a<br />

common food resource <strong>for</strong> owls regardless of<br />

breeding status. These different results may<br />

reflect variation in prey availability, sampling<br />

bias and variation, temporal variation, or true<br />

regional differences in owl diets.<br />

To evaluate if any particular prey group<br />

could increase persistence of <strong>the</strong> <strong>Mexican</strong><br />

spotted owl, we analyzed owl reproductive<br />

success (average number of young fledged) as a<br />

function of diet, year, and geographic location<br />

using a three-factor ANOVA. This approach<br />

would allow us to quantify variation in owl<br />

reproduction that could be attributed to <strong>the</strong><br />

frequency of a particular prey in <strong>the</strong> owl’s diet.<br />

<strong>Owl</strong> reproduction related to consumption<br />

frequency of a given prey group would imply a<br />

relative importance of that prey. In using this<br />

approach, we fur<strong>the</strong>r assumed that prey species<br />

that were positively related to reproduction also<br />

enhanced <strong>the</strong> owl’s survival.


14<br />

Table Table 5.8. 5.8. Factors influencing trends observed in <strong>Mexican</strong> spotted owl diets. Data are from three studies conducted in nor<strong>the</strong>rn Arizona, <strong>the</strong> Sacramento<br />

Mountains, New Mexico, and <strong>the</strong> Tularosa Mountains, New Mexico, during breeding seasons of 1991, 1992, and 1993.<br />

FPO


15<br />

Table Table 5.8. 5.8. (continued)<br />

FPO


This analysis was conducted by stratifying<br />

and comparing variation in <strong>the</strong> number of<br />

young owls produced among three study areas<br />

(SACMT2, COCODM, GILADM), three<br />

different breeding seasons (1991-1993), and<br />

three amounts of prey consumption, low<br />

(< 33%), high (³ 66%), or medium. Seven prey<br />

groups were examined in separate ANOVAs to<br />

ensure independence among diet frequencies.<br />

These included woodrats, peromyscid mice,<br />

voles, gophers, rabbits, o<strong>the</strong>r medium-sized<br />

mammals, and arthropods. Study area and year<br />

were included as blocking factors to account <strong>for</strong><br />

spatial and temporal effects that might alternatively<br />

explain patterns in <strong>the</strong> owl’s reproduction,<br />

respectively. This analysis differed from <strong>the</strong> 3factor<br />

ANOVA used to examine geographic<br />

variation in <strong>the</strong> owl’s diet where number of<br />

young was treated as a predictor variable ra<strong>the</strong>r<br />

than a response.<br />

Results from each ANOVA suggested that<br />

<strong>the</strong> owl’s reproductive success was not influenced<br />

by a single prey species but ra<strong>the</strong>r by many<br />

species in combination. None of <strong>the</strong> specific<br />

prey groups significantly influenced owl reproductive<br />

success when diet frequencies were<br />

examined separately (Table 5.9). Time was a<br />

significant factor influencing <strong>the</strong> owl’s reproduction<br />

(Table 5.9) when consumption of woodrats<br />

or voles was examined. Prey abundance is known<br />

to fluctuate through time. Thus, effects attributed<br />

to temporal variation may also represent<br />

associated changes in prey populations. However,<br />

it is more likely that <strong>the</strong> owl’s reproductive<br />

success was influenced by total prey biomass<br />

consumed in a given year, ra<strong>the</strong>r than by a single<br />

prey species. For example, when frequencies of<br />

<strong>the</strong> three most common prey groups, woodrats,<br />

peromyscid mice and voles were combined and<br />

analyzed, diet was a nearly significant predictor<br />

of <strong>the</strong> owl’s reproductive success (F = 2.930,<br />

d.f. = 2, 98, P = 0.058). More owl young were<br />

produced when moderate to high amounts of<br />

<strong>the</strong>se common prey were consumed in all three<br />

study areas (Figure 5.3).<br />

Contrary to <strong>the</strong> above results, certain prey<br />

species may be more important in certain<br />

regions of <strong>the</strong> owl’s range. For example, o<strong>the</strong>r<br />

in<strong>for</strong>mation from Ward et al. (unpublished)<br />

suggests that reproductive success of <strong>Mexican</strong><br />

Volume II/Chapter 5<br />

16<br />

spotted owls in <strong>the</strong> Sacramento Mountains may<br />

increase when deer mice populations irrupt. In<br />

1991, deer mouse biomass averaged 0.911 kg/ha<br />

in mixed-conifer <strong>for</strong>ests (Figure 5.4a) and <strong>the</strong><br />

number of young produced corresponded to <strong>the</strong><br />

number of peromyscid mice consumed (Figure<br />

5.4b). In 1992 and 1993, owl reproductive<br />

success decreased corresponding with a reduction<br />

in deer mouse abundance and <strong>the</strong> frequency of<br />

peromyscid mice consumed by owls (Figure<br />

5.4c). Reproduction among owls dwelling in<br />

steep-walled canyons of <strong>the</strong> Colorado Plateau<br />

may also depend on a specialized diet because<br />

<strong>the</strong>se owls consume a greater number of<br />

woodrats compared to o<strong>the</strong>r localities (Table<br />

5.1). These results may be exceptions, as most of<br />

our findings support maintenance of several<br />

common prey species, ra<strong>the</strong>r than enhancing<br />

populations of a few.<br />

ZOOGEOGRAPHY ZOOGEOGRAPHY AND<br />

AND<br />

MACROHABITATS<br />

MACROHABITATS<br />

OF OF COMMON COMMON PREY<br />

PREY<br />

As detailed earlier, <strong>Mexican</strong> spotted owls use<br />

a variety of prey. Species regarded as “common”<br />

are those comprising ³10% of <strong>the</strong> owl diet by<br />

relative frequency or biomass within a given RU.<br />

Prey commonly consumed by <strong>the</strong> owl varies<br />

geographically (Table 5.10). This geographic<br />

variation can be attributed to two primary<br />

factors: <strong>the</strong> geographic range of <strong>the</strong> prey and <strong>the</strong><br />

degree of sympatry in macrohabitats of <strong>the</strong> owl<br />

and its prey. Below, we provide an overview of<br />

basic macrohabitat associations of common owl<br />

prey.<br />

Bats Bats<br />

Bats<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

MAMMALS<br />

MAMMALS<br />

As a group, bats are common prey in portions<br />

of <strong>the</strong> Colorado Plateau, Upper Gila<br />

Mountains, and Basin and Range - West RUs<br />

and <strong>the</strong>y are consumed occasionally by owls in<br />

all RUs. If taken from outside roosts or nursery<br />

colonies, bats may represent a low-cost, opportunistic<br />

food source <strong>for</strong> <strong>the</strong> owls. No particular<br />

species appears to be used in great abundance;<br />

but when considered as a group, <strong>the</strong> importance


17<br />

Table Table 5.9. 5.9. Factors influencing production of <strong>Mexican</strong> spotted owls in nor<strong>the</strong>rn Arizona, <strong>the</strong> Sacramento Mountains, New Mexico, and <strong>the</strong> Tularosa<br />

Mountains, New Mexico, during 1991, 1992, and 1993. Consumption of common prey was categorized according to <strong>the</strong> frequency of that prey in <strong>the</strong><br />

owls' diet; low ( 0.05) in tests on<br />

peromyscid mice or could not be quantified in tests with o<strong>the</strong>r prey because of limited sample sizes.<br />

FPO


18<br />

Table Table 5.9. 5.9. (continued)<br />

FPO


Figure Figure 5.3. 5.3. Reproductive success (mean number of young produced) as a function of low (


Figure Figure 5.4. 5.4. Biomass (kg/ha) of (a) (a) common prey occurring in mixed-conifer <strong>for</strong>ests, (b) (b) frequencies<br />

of peromyscid mice consumed by <strong>Mexican</strong> spotted owls stratified by number of owl young produced,<br />

and (c) (c) average (±95% CI) number of owl young produced in <strong>the</strong> Sacramento Mountains, New<br />

Mexico (Ward et al. unpublished). Bars in (b) (b) represent variation among owl pairs (standard errors).<br />

Volume II/Chapter 5<br />

20<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong>


21<br />

Table Table Table 5.10. 5.10. 5.10. Prey comprising ³10% of relative frequency (X) or biomass (O) in <strong>the</strong> diet of <strong>Mexican</strong> spotted owls.<br />

FPO


occur within woodlands and <strong>for</strong>ests. Botta’s<br />

pocket gopher, nor<strong>the</strong>rn pocket gopher, and<br />

sou<strong>the</strong>rn pocket gopher occur within <strong>the</strong> range<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl. Botta’s pocket<br />

gopher is <strong>the</strong> most widespread, being found in<br />

most vegetation types. The nor<strong>the</strong>rn pocket<br />

gopher occurs in montane habitats of <strong>the</strong> nor<strong>the</strong>rn<br />

part of <strong>the</strong> owls’ range. The sou<strong>the</strong>rn pocket<br />

gopher is found in Basin and Range - West and<br />

<strong>Mexican</strong> RUs. Both Hoffmeister (1986) and<br />

Findley et al. (1975) noted that Botta’s pocket<br />

gopher is ubiquitous whenever soil is suitable <strong>for</strong><br />

burrowing, whereas <strong>the</strong> nor<strong>the</strong>rn pocket gopher<br />

is more typically found in parks and meadows<br />

within montane <strong>for</strong>ests. Findley et al. (1975)<br />

stated that <strong>the</strong> sou<strong>the</strong>rn pocket gopher inhabits<br />

shallow, rocky soils of pine <strong>for</strong>ests.<br />

Peromyscid Peromyscid Mice Mice<br />

Mice<br />

Eight peromyscid mice occur within <strong>the</strong><br />

range of <strong>the</strong> <strong>Mexican</strong> spotted owl. Only two<br />

species, <strong>the</strong> deer mouse and brush mouse are<br />

consumed regularly by owls in all RUs (Table<br />

5.10). A third species, <strong>the</strong> canyon mouse, is<br />

likely consumed by owls dwelling in <strong>the</strong> Colorado<br />

Plateau RU and <strong>the</strong> rock mouse has been<br />

reported in <strong>the</strong> diet of owls occurring in <strong>the</strong><br />

Basin and Range - East RU (Table 5.7).<br />

The deer mouse is widespread, inhabiting all<br />

vegetation types except high-elevation tundra<br />

(Bailey 1931, Hall and Kelson 1959, Armstrong<br />

1977, Goodwin and Hunger<strong>for</strong>d 1979). High<br />

reproductive success of spotted owls in <strong>the</strong><br />

Sacramento Mountains, New Mexico, (Basin<br />

and Range - East RU) has been recorded during<br />

irruptions of deer mice in mixed-conifer <strong>for</strong>ests<br />

(See Abundance and Distribution of Common<br />

Prey, Sacramento Mountains).<br />

More restricted in distribution, <strong>the</strong> brush<br />

mouse typically inhabits areas with extensive<br />

rock and shrub cover in pinyon-juniper, riparian,<br />

oak, and pine-oak woodlands (Wilson 1968,<br />

Armstrong 1979, Svoboda et al. 1988).<br />

Goodwin and Hunger<strong>for</strong>d (1979) found that<br />

brush mice inhabit rocky slopes in central<br />

Arizona’s pine-oak <strong>for</strong>ests, but rarely use pure<br />

stands of pine.<br />

The canyon mouse occupies <strong>the</strong> canyon<br />

walls, cliffs, and steep rocky slopes of nor<strong>the</strong>rn<br />

Volume II/Chapter 5<br />

22<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Arizona, Utah, northwestern New Mexico, and<br />

western Colorado (Armstrong 1979, Johnson<br />

and Armstrong 1987). Findley et al. (1975)<br />

reported that <strong>the</strong> canyon mouse is often found<br />

in pinyon-juniper woodlands at <strong>the</strong> base of cliffs,<br />

although Johnson and Armstrong (1987) noted<br />

that vegetation associations are of limited importance<br />

relative to <strong>the</strong> presence of suitable rocky<br />

substrates.<br />

The rock mouse is found in association with<br />

rocky substrates generally above 1,900 m (6,230<br />

ft) elevation (Findley et al. 1975, Cornely et al.<br />

1981). This species occurs in most of <strong>the</strong> United<br />

States portion of <strong>the</strong> owl’s range and is often<br />

sympatric with deer mice and brush mice (Wilson<br />

1968, Cornely et al. 1981, Ribble and<br />

Samson 1987).<br />

Woodrats Woodrats<br />

Woodrats<br />

<strong>Mexican</strong>, bushy-tailed, desert, and whitethroated<br />

woodrats are consumed by <strong>Mexican</strong><br />

spotted owls. For <strong>the</strong> owl, woodrats provide a<br />

large mass of food per capture. This prey is a<br />

primary food source <strong>for</strong> owls through most of its<br />

range but particularly in <strong>the</strong> canyon habitats of<br />

<strong>the</strong> Colorado Plateau RU, where all four species<br />

of woodrats are sometimes found (Table 5.10).<br />

The <strong>Mexican</strong> woodrat is perhaps <strong>the</strong> most<br />

common woodrat found within <strong>the</strong> range of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl. It occurs within all RUs,<br />

although populations are disjunct because of <strong>the</strong><br />

species’ montane distribution (Cornely and<br />

Baker 1986). The altitudinal range of <strong>the</strong> <strong>Mexican</strong><br />

woodrat begins in <strong>the</strong> lower pine zone and<br />

extends upward through mixed-conifer <strong>for</strong>ests<br />

where Findley et al. (1975) reported <strong>the</strong>y reach<br />

<strong>the</strong>ir greatest abundance. Hoffmeister (1986; see<br />

also Goodwin and Hunger<strong>for</strong>d 1979) regarded<br />

<strong>Mexican</strong> woodrats as rock dwellers within <strong>the</strong>se<br />

vegetation types, infrequently extending into<br />

pinyon-juniper woodlands. Armstrong (1972),<br />

however, reported that this species typically uses<br />

pinyon-juniper woodland in western Colorado<br />

and scrub-like oaks and mountain mahogany<br />

vegetation along <strong>the</strong> eastern foothills northwards<br />

almost to Wyoming. The range of <strong>Mexican</strong><br />

woodrats in Utah is restricted to <strong>the</strong> sou<strong>the</strong>astern<br />

portion of <strong>the</strong> state, east of <strong>the</strong> Colorado River.


Presumably, <strong>the</strong> species uses similar habitats in<br />

Utah as it does in western Colorado.<br />

Bushy-tailed woodrats are a common diet<br />

component within <strong>the</strong> Colorado Plateau and<br />

both Sou<strong>the</strong>rn Rocky Mountain RUs. This<br />

species is a cordilleran mammal of western<br />

Colorado, northcentral and northwestern New<br />

Mexico, nor<strong>the</strong>astern Arizona, and eastern Utah<br />

(Durrant 1952, Hoffmeister 1986, Findley et al.<br />

1975, Armstrong 1972, 1979). Finley (1958)<br />

reported that bushy-tailed woodrats use a variety<br />

of vegetation types, primarily woodland and<br />

shrublands. Their distribution may depend on<br />

<strong>the</strong> presence of suitable rock outcrops ra<strong>the</strong>r<br />

than specific vegetation (Finley 1958,<br />

Hoffmeister 1986). At higher elevations, <strong>the</strong>se<br />

outcrops interrupt open <strong>for</strong>ests of Douglas-fir,<br />

aspen, or ponderosa pine containing a welldeveloped<br />

shrub understory. Typically, lower<br />

elevation sites are dominated by pinyon and<br />

juniper.<br />

The desert woodrat is consumed by owls in<br />

<strong>the</strong> Colorado Plateau RU. This species is most<br />

abundant in <strong>the</strong> Arizona strip where it inhabits a<br />

variety of plant communities including creosote<br />

bush and cacti to junipers and pine (Hoffmeister<br />

1986). Desert woodrats frequently nest in<br />

crevices of cliffs and rock outcrops within<br />

juniper or shadscale plant communities of Utah<br />

and Colorado, below 2,000 m (6,560 ft) elevation<br />

(Finley 1958).<br />

White-throated woodrats are common prey<br />

of owls occurring in <strong>the</strong> Upper Gila Mountains<br />

and Basin and Range - West RUs and are less<br />

frequently consumed by owls in o<strong>the</strong>r RUs. This<br />

woodrat species is typically distributed below <strong>the</strong><br />

conifer belt, although it can be found in pinyonjuniper<br />

woodlands (Hoffmeister 1986).<br />

Voles<br />

Voles<br />

Four species of voles are common prey of <strong>the</strong><br />

<strong>Mexican</strong> spotted owl including <strong>the</strong> <strong>Mexican</strong> (or<br />

Mogollon vole [after Frey and LaRue 1993]),<br />

mountain, meadow, and long-tailed voles (Table<br />

5.10). Three of <strong>the</strong>se species can be ordered<br />

along an environmental moisture gradient from<br />

semi-arid (<strong>Mexican</strong> vole), mesic (mountain<br />

vole), to hydric (meadow vole). Long-tailed voles<br />

Volume II/Chapter 5<br />

23<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

inhabit environments along <strong>the</strong> entire gradient<br />

(Getz 1985).<br />

The <strong>Mexican</strong> vole is common within <strong>the</strong><br />

greatest number of RUs, including Basin and<br />

Range - East, Sou<strong>the</strong>rn Rocky Mountains - New<br />

Mexico, and Upper Gila Mountains RUs. It is<br />

fairly widely distributed in Arizona and New<br />

Mexico, but it is confined to <strong>the</strong> sou<strong>the</strong>ast part<br />

of Utah and to southwest Colorado. This<br />

species occurs in <strong>the</strong> widest range of habitats of<br />

any microtine and is generally associated with<br />

xeric grassy locations extending from pinyonjuniper<br />

to spruce-fir zones (Armstrong 1972,<br />

Findley and Jones 1962, Findley et al. 1975,<br />

Finley et al. 1986, Hoffmeister 1986, Frey and<br />

LaRue 1993).<br />

The two species associated with wet conditions<br />

(mountain and meadow voles) generally<br />

occur in <strong>the</strong> nor<strong>the</strong>rn portion of <strong>the</strong> owl’s range.<br />

Mountain voles are common prey in <strong>the</strong> Colorado<br />

Plateau and both Sou<strong>the</strong>rn Rocky Mountain<br />

RUs. In <strong>the</strong>se areas, <strong>the</strong> mountain vole<br />

occupies <strong>for</strong>est meadows ranging in elevation<br />

from open pine-oak to spruce-fir <strong>for</strong>ests.<br />

Armstrong (1977) found mountain voles in<br />

dense grass cover with a sparse overstory. This<br />

vole’s geographic range includes most montane<br />

regions along <strong>the</strong> north-south axes of Colorado<br />

and Utah, nor<strong>the</strong>rn New Mexico, and <strong>the</strong> White<br />

Mountains in Arizona. Meadow voles occur in<br />

both Sou<strong>the</strong>rn Rocky Mountains RUs where<br />

permanent water is provided by springs and<br />

marshes (Armstrong 1972, Findley et al. 1975,<br />

Finley et al. 1986).<br />

The long-tailed vole occurs within <strong>the</strong><br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico,<br />

Upper Gila Mountains, and Basin and Range -<br />

East RUs. Findley et al. (1975) reported that it is<br />

associated with meadows and <strong>for</strong>est edge, being<br />

most common in <strong>the</strong> mixed-conifer and sprucefir<br />

zones but also using mixed-conifer stringers<br />

found along canyons in <strong>the</strong> ponderosa pine<br />

zone. Armstrong (1972, 1977) noted that this<br />

microtine requires a grassy understory less than<br />

o<strong>the</strong>r vole species because it is often found<br />

within <strong>for</strong>ests supporting minimal grass cover.


Birds<br />

Birds<br />

Birds are common prey <strong>for</strong> <strong>the</strong> owls in <strong>the</strong><br />

Sou<strong>the</strong>rn Rocky Mountains - New Mexico,<br />

Upper Gila Mountains, and Basin and<br />

Range - West RUs where numerous species have<br />

been identified in pellets (Table 5.7). The<br />

importance of birds to spotted owls is uncertain.<br />

Birds do contribute to <strong>the</strong> diversity of prey taken<br />

by owls and may provide food resources when<br />

small mammals are less abundant. However, use<br />

of birds as prey is likely seasonal because many of<br />

<strong>the</strong> passerine species consumed by owls are<br />

migratory. All species identified in <strong>the</strong> owls’ diet<br />

to date are <strong>for</strong>est dwellers.<br />

Arthropods<br />

Arthropods<br />

Arthropods are common prey of spotted<br />

owls in <strong>the</strong> Colorado Plateau, Sou<strong>the</strong>rn Rocky<br />

Mountains - Colorado, Sou<strong>the</strong>rn Rocky Mountains<br />

- New Mexico, and Upper Gila Mountains<br />

RUs. Many species are consumed (Table 5.7).<br />

The importance of arthropods to spotted owls is<br />

also uncertain. If one considers biomass alone,<br />

arthropods contribute little to <strong>the</strong> owl’s diet.<br />

However, arthropods may provide a low cost,<br />

high quality food taken opportunistically.<br />

Description of macrohabitats of arthropods<br />

consumed by spotted owls is beyond <strong>the</strong> scope<br />

of this document. Active management of arthropod<br />

populations <strong>for</strong> owl recovery is probably not<br />

necessary.<br />

Volume II/Chapter 5<br />

ABUNDANCE ABUNDANCE ABUNDANCE AND AND<br />

AND<br />

DISTRIBUTION DISTRIBUTION OF OF<br />

OF<br />

COMMON COMMON PREY<br />

PREY<br />

The availability of prey to <strong>Mexican</strong> spotted<br />

owls depends on prey abundance, <strong>the</strong> vulnerability<br />

of <strong>the</strong> prey to capture by <strong>the</strong> owl, and <strong>the</strong><br />

probability that <strong>the</strong> owl and its prey occur in <strong>the</strong><br />

same habitat. All of <strong>the</strong>se factors vary by habitat<br />

condition. In addition, <strong>the</strong> amount of energy<br />

available to <strong>the</strong> owl will vary according to <strong>the</strong><br />

type, size, and condition of <strong>the</strong> prey. Thus, it is<br />

useful to convert prey numbers into values that<br />

reflect energy input. Because rodents are <strong>the</strong><br />

most common prey of <strong>the</strong> spotted owl, we<br />

24<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

assume that most of <strong>the</strong> difference in energy<br />

content among mammalian prey can be attributed<br />

to body mass.<br />

Although no in<strong>for</strong>mation exists on <strong>the</strong><br />

vulnerability of different prey species to capture<br />

by <strong>Mexican</strong> spotted owls, estimates of common<br />

prey abundance and mass within several different<br />

vegetation communities are available from<br />

research conducted in nor<strong>the</strong>rn Arizona (Block<br />

and Ganey, unpublished) and <strong>the</strong> Sacramento<br />

Mountains, New Mexico (Ward et al., unpublished).<br />

These two studies provide estimates of<br />

biomass and microhabitat associations (See Prey<br />

Habitat) <strong>for</strong> <strong>the</strong> owl’s common prey in four<br />

different vegetation communities used by owls<br />

<strong>for</strong> <strong>for</strong>aging, ponderosa pine-Gambel oak,<br />

mixed-conifer, high-elevation meadows, and<br />

ponderosa pine-pinyon-juniper woodlands. The<br />

methods and results of both studies are briefly<br />

described below.<br />

Nor<strong>the</strong>rn Nor<strong>the</strong>rn Arizona<br />

Arizona<br />

Methods<br />

Methods<br />

Small mammal populations were sampled<br />

using live-trapping and mark-recapture techniques<br />

from November 1990 through December<br />

1992 within home ranges of five owl pairs. The<br />

purpose of <strong>the</strong> sampling was to estimate biomass<br />

of <strong>the</strong> owl’s common prey and determine <strong>the</strong><br />

prey’s distribution. Common prey were determined<br />

from owl pellets collected during this<br />

study.<br />

The study area consisted of ponderosa pine-<br />

Gambel oak <strong>for</strong>est, although each trapping grid<br />

was unique with respect to <strong>the</strong> relative composition<br />

and structure of <strong>the</strong> vegetation. Ancillary<br />

trapping was conducted within <strong>the</strong> winter range<br />

of two owls that migrated downward to pinyonjuniper<br />

woodland in <strong>the</strong> Verde Valley. This<br />

trapping was confined to five sets of smaller<br />

trapping grids (described below) <strong>for</strong> a total of<br />

574 trap nights.<br />

Trapping grids were randomly established at<br />

general <strong>for</strong>aging areas identified by radio telemetry<br />

(Ganey and Block, unpublished data). Traps<br />

were arrayed in 10 x 10 or 2 x 10 grids with<br />

20-m (65.5 ft) spacings between stations. The


larger grids were used to estimate both density<br />

and habitat correlates; <strong>the</strong> smaller grids were<br />

used only to assess habitat use by species (see<br />

Prey Habitat). Large Sherman live traps (size<br />

8 x 9 x 23 cm [3 x 3.5 x 9 in]) were placed at<br />

each grid station; extra-large Sherman live traps<br />

(size 10 x 18 x 60 cm (4 x 4.5 x 15 in) were<br />

placed at alternate stations. Two sizes of traps<br />

were used to minimize <strong>the</strong> potential bias against<br />

capturing larger prey such as woodrats in <strong>the</strong><br />

smaller traps.<br />

Grids were trapped from 3-7 nights during<br />

each trapping session. Traps were left open<br />

during <strong>the</strong> day to sample diurnal sciurids. The<br />

goal was to continue trapping until 90% of all<br />

captures were recaptures to approach assumptions<br />

of population closure. This goal was<br />

typically reached between five and seven nights.<br />

During periods of inclement wea<strong>the</strong>r, however,<br />

we relaxed this goal to minimize trapping<br />

mortalities. Each grid was trapped <strong>for</strong> 5-7<br />

sessions during <strong>the</strong> study. Total trapping ef<strong>for</strong>t<br />

was 49,911 trapnights (adjusted <strong>for</strong> closed and<br />

unoccupied, or o<strong>the</strong>rwise unavailable traps).<br />

When an animal was captured, it was identified<br />

to species, weighed, and marked by toe clipping.<br />

Abundance was estimated as <strong>the</strong> number of<br />

individuals per effective sampling area (ha) using<br />

closed population estimators (Program CAP-<br />

TURE; Otis et al. 1978, White et al. 1982).<br />

Although <strong>the</strong> number of individuals captured at<br />

some grids during some seasons was insufficient<br />

<strong>for</strong> producing unbiased estimates, we considered<br />

<strong>the</strong> closed population estimators in Program<br />

CAPTURE more appropriate than minimum<br />

numbers alive per grid area. The <strong>for</strong>mer permitted<br />

estimating capture probabilities and sampling<br />

variances plus population size; <strong>the</strong> latter<br />

would not.<br />

Biomass of common prey, expressed as<br />

kilograms per hectare (kg/ha), was calculated as a<br />

product of prey density and average mass. The<br />

delta method (Goodman 1960) was used to<br />

estimate <strong>the</strong> sampling variance of biomass from<br />

<strong>the</strong> variances associated with sampling prey<br />

density and mass.<br />

Volume II/Chapter 5<br />

25<br />

General General Distribution Distribution<br />

Distribution<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

The three primary prey species captured in<br />

ponderosa pine-Gambel oak <strong>for</strong>ests were deer<br />

mouse, brush mouse, and <strong>Mexican</strong> woodrat.<br />

O<strong>the</strong>r prey species captured included pinyon<br />

mouse, white-throated woodrat, Stephens’<br />

woodrat, <strong>Mexican</strong> vole, rock squirrel, graycollared<br />

chipmunk, and cliff chipmunk. Brush<br />

mice and white-throated woodrats were captured<br />

with <strong>the</strong> greatest relative frequency in pinyonjuniper<br />

woodlands. The trapping ef<strong>for</strong>ts did not<br />

sample prey species such as pocket gophers,<br />

cottontail rabbits, birds, and arthropods. Fur<strong>the</strong>r,<br />

only <strong>the</strong> three primary species were captured<br />

in sufficient numbers to permit density<br />

calculations or estimates of habitat correlates.<br />

Thus, our analyses of population size and habitat<br />

use address only <strong>the</strong>se three species.<br />

Population Population Abundance<br />

Abundance<br />

The deer mouse was <strong>the</strong> most abundant<br />

species captured, followed by <strong>the</strong> brush mouse<br />

and <strong>Mexican</strong> woodrat. A seasonal trend of<br />

decreased prey biomass during winter was noted<br />

(Figure 5.5a). Prey abundance and biomass also<br />

varied by year (Figure 5.5a) and among owl<br />

territories (Block and Ganey, unpublished data).<br />

Sacramento Sacramento Mountains<br />

Mountains<br />

Methods<br />

Methods<br />

The Sacramento Mountains are located in<br />

southcentral New Mexico. An investigation of<br />

<strong>the</strong> abundance and distribution of common<br />

mammalian prey began in 1991 and is ongoing<br />

(Ward et al., unpublished data). Common prey<br />

were determined from owl pellets collected<br />

during this study. Abundance was estimated<br />

among three general vegetation communities<br />

using mark-recapture methodology.<br />

The three communities included mesic<br />

<strong>for</strong>ests, xeric <strong>for</strong>ests, and meadows. Mesic <strong>for</strong>ests<br />

were a mixture of Douglas-fir, white fir, southwestern<br />

white pine, ponderosa pine, and<br />

Englemann spruce (i.e., mixed conifer). Meadows<br />

consisted of <strong>for</strong>bs and grasses and were


Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure Figure 5.5. 5.5. Average biomass of common prey (kg/ha) of <strong>the</strong> <strong>Mexican</strong> spotted owl in (a) (a) pine-<br />

Gambel oak habitats of nor<strong>the</strong>rn Arizona (Block and Ganey unpublished data) and (b) (b) in three habitats<br />

of <strong>the</strong> Sacramento Mountains, New Mexico. Vertical bars are standard errors. Lower horizontal<br />

bars in (b) (b) (b) separate sampling error from errors associated with variation among sites. Number of sites<br />

within each habitat are shown in paren<strong>the</strong>sis.<br />

26


associated with drainage bottoms adjacent to<br />

mesic <strong>for</strong>ests. Xeric <strong>for</strong>ests included ponderosa<br />

pine, pinyon, juniper, and low-growing oaks.<br />

The same two sizes of live-traps described <strong>for</strong><br />

<strong>the</strong> nor<strong>the</strong>rn Arizona study were used to capture<br />

prey in <strong>the</strong> Sacramento Mountains. Traps were<br />

also arranged similarly, with <strong>the</strong> following<br />

exceptions. In 1991, four 13-ha (32.1-ac)<br />

trapping grids were established as part of a pilot<br />

study. Two grids were placed in <strong>the</strong> mesic <strong>for</strong>est<br />

and two in <strong>the</strong> xeric <strong>for</strong>est. In 1992 and subsequent<br />

years, six 4-ha (9.9-ac) trapping grids were<br />

placed in each of <strong>the</strong> mesic and xeric <strong>for</strong>est<br />

types. Traps were arrayed as 11 x 11 station grids.<br />

In addition, six 1.8-ha (4.5-ac) grids were<br />

established in meadows. Each of <strong>the</strong> latter grids<br />

consisted of 105 large Sherman traps spaced<br />

15 m (49.2 ft) apart in a 5 x 21 array.<br />

All grid sites were selected randomly from a<br />

list of known <strong>Mexican</strong> spotted owl territories.<br />

Grids were placed as close to a nest or roost area<br />

as possible while maintaining homogeneity at a<br />

scale analogous to a <strong>for</strong>est stand. In addition,<br />

each grid was £ 0.8 km (0.5 mi) from a nest or<br />

roost to ensure that <strong>the</strong> site was available <strong>for</strong> use<br />

by spotted owls.<br />

All captured individuals were marked with<br />

uniquely numbered ear tags in both ears. Loss of<br />

both tags during an eight-night trapping session<br />

was less than 1% <strong>for</strong> all marked species. All o<strong>the</strong>r<br />

methods of sampling and data collection were<br />

similar to <strong>the</strong> study in nor<strong>the</strong>rn Arizona (Block<br />

and Ganey, unpublished data). We <strong>the</strong>re<strong>for</strong>e<br />

considered <strong>the</strong> results of both studies to be<br />

comparable. Prey abundance and biomass were<br />

estimated using <strong>the</strong> same procedures described<br />

<strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn Arizona study. Tests <strong>for</strong> spatial<br />

and temporal differences in prey biomass were<br />

conducted using a two-factor ANOVA.<br />

General General Distribution<br />

Distribution<br />

Common prey included woodrats,<br />

peromyscid mice, voles, arthropods, cottontail<br />

rabbits, and o<strong>the</strong>r medium-sized mammals such<br />

as long-tailed weasels, red squirrels, and chipmunks<br />

(Tables 5.3 and 5.6). Cottontails and<br />

o<strong>the</strong>r medium-sized mammals were consumed<br />

infrequently but larger estimates of mass from a<br />

few individuals resulted in larger biomass calcu-<br />

Volume II/Chapter 5<br />

27<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

lations from <strong>the</strong>se prey groups. However, <strong>the</strong><br />

exact sizes of consumed individuals were undetermined<br />

and <strong>the</strong> range of mass <strong>for</strong> larger prey<br />

consumed by owls could have been considerable.<br />

For example, <strong>the</strong> size of a cottontail taken by a<br />

spotted owl could range 50 to 400 g (1.8-14.1<br />

oz) whereas a peromyscid mouse could range 10<br />

to 40 g (0.4 to 1.4 oz). This potentially results in<br />

upwardly biased estimates of biomass <strong>for</strong> larger<br />

species in <strong>the</strong> owls’ diet. For this reason, we are<br />

uncertain about <strong>the</strong> role of cottontails and o<strong>the</strong>r<br />

medium-sized prey as common food resources<br />

<strong>for</strong> <strong>the</strong> owl. In contrast, arthropods comprised<br />

11.4% of <strong>the</strong> diet but were not considered<br />

energetically important because of <strong>the</strong>ir small<br />

mass (1-2 g [0.04-0.07 oz]).<br />

Five species common in <strong>the</strong> owl’s diet were<br />

captured regularly during live-trapping. The<br />

distribution of each species varied by vegetation<br />

community. Deer mice were found in all three<br />

communities. Brush mice were restricted to <strong>the</strong><br />

xeric <strong>for</strong>est type and were primarily associated<br />

with areas containing shrub-<strong>for</strong>m oaks. Longtailed<br />

voles and <strong>Mexican</strong> voles were more common<br />

in meadows, but <strong>the</strong>y also occupied <strong>the</strong><br />

transition zones between meadows and mesic<br />

<strong>for</strong>ests. <strong>Mexican</strong> voles were also found infrequently<br />

in xeric <strong>for</strong>ests. <strong>Mexican</strong> woodrats<br />

occupied mesic <strong>for</strong>ests and <strong>the</strong> ecotone between<br />

<strong>the</strong>se <strong>for</strong>ests and meadows, as well as xeric<br />

<strong>for</strong>ests. Medium-sized mammals such as <strong>the</strong><br />

gray-footed chipmunk were found in mesic<br />

<strong>for</strong>ests, <strong>the</strong> edges between mesic <strong>for</strong>ests and<br />

meadows, and rarely in xeric <strong>for</strong>ests. Cottontail<br />

rabbits and red squirrels were only encountered<br />

in mesic <strong>for</strong>ests. However, our sampling procedures<br />

were inadequate <strong>for</strong> estimating abundance<br />

and distribution of <strong>the</strong>se species and of<br />

arthropods.<br />

Population Population Abundance Abundance<br />

Abundance<br />

Total biomass (kg/ha) of five common prey<br />

(Figure 5.5b) varied annually over <strong>the</strong> three<br />

summers 1991-1993 (F = 12.7, d.f. = 2, 32,<br />

P < 0.001) and also seasonally during <strong>the</strong> summer-fall-winter<br />

period of 1993-1994 (F = 17.0,<br />

d.f. = 2, 21, P < 0.001). Fluctuations in prey<br />

biomass also varied by vegetation community<br />

during both annual (F = 5.5, d.f. = 2, 32,


P = 0.009) and seasonal cycles (F = 5.8, d.f. = 2,<br />

21, P = 0.010). The general trend was that prey<br />

biomass was moderately high during <strong>the</strong> summer<br />

(when owls feed <strong>the</strong>ir young) of 1991, peaking<br />

in 1992, <strong>the</strong>n decreasing to moderate levels in<br />

1993 (Figure 5.5b). Fur<strong>the</strong>r, temporal trends<br />

differed by vegetation community. Prey biomass<br />

decreased in mesic <strong>for</strong>ests and meadows over<br />

consecutive summers but increased <strong>the</strong>n decreased<br />

in xeric <strong>for</strong>ests during this same period<br />

(Figure 5.5b). This relationship was evident by a<br />

statistically significant interaction of vegetation<br />

community and year on prey biomass (F = 3.3,<br />

d.f. = 3, 32, P = 0.032). Whereas prey biomass<br />

pooled across all communities decreased from<br />

<strong>the</strong> summer to fall of 1993 be<strong>for</strong>e increasing<br />

slightly in <strong>the</strong> winter (Figure 5.5b), seasonal<br />

trends differed among communities (F = 17.0,<br />

d.f. = 2, 21, P < 0.001). These patterns also<br />

showed an interactive influence of season and<br />

vegetation community on prey biomass (F = 3.0,<br />

d.f. = 2, 21, P = 0.044). Thus, abundance of<br />

potential food resources <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> spotted<br />

owl in <strong>the</strong> Sacramento Mountains are temporally<br />

variable and habitat dependent.<br />

The variations in total prey biomass were<br />

also reflected in <strong>the</strong> population dynamics of<br />

different species (Figure 5.6a). For example, deer<br />

mice densities within mesic <strong>for</strong>ests peaked in <strong>the</strong><br />

summer of 1991 and declined through <strong>the</strong><br />

summer of 1993 while maintaining low densities<br />

in <strong>the</strong> meadows (Figure 5.6b). Deer mice in<br />

xeric <strong>for</strong>ests maintained lower, relatively stable<br />

densities that peaked in summer of 1992 (Figure<br />

5.6c). In contrast, <strong>Mexican</strong> voles maintained<br />

low, stable populations in <strong>the</strong> mesic <strong>for</strong>ests but<br />

increased dramatically in meadows and moderately<br />

in xeric <strong>for</strong>ests in <strong>the</strong> summer of 1992<br />

be<strong>for</strong>e crashing in both habitats during <strong>the</strong><br />

summer of 1993 (Figure 5.6).<br />

In summary, <strong>the</strong> two studies indicate that<br />

<strong>the</strong> owl’s food resources are quite variable among<br />

vegetation communities and through time.<br />

Arranging <strong>the</strong> four vegetation communities<br />

examined in <strong>the</strong>se two studies in descending<br />

amount of summer prey biomass indicates that<br />

meadows > mixed-conifer <strong>for</strong>est > ponderosa<br />

pine-pinyon-juniper-oak woodlands > ponderosa<br />

pine-Gambel oak <strong>for</strong>est. When considering<br />

o<strong>the</strong>r factors that influence <strong>the</strong> availability of<br />

Volume II/Chapter 5<br />

28<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

prey, mixed-conifer <strong>for</strong>ests likely provide <strong>the</strong><br />

greatest amount of food during summer periods.<br />

Rearranging <strong>the</strong> same communities according to<br />

winter prey biomass indicates that meadows ><br />

ponderosa pine-pinyon-juniper-oak woodlands ><br />

ponderosa pine-Gambel oak <strong>for</strong>est > mixedconifer<br />

<strong>for</strong>est. Accounting <strong>for</strong> <strong>the</strong> availability of<br />

prey, woodlands with a mixture of ponderosa<br />

pine, pinyon-juniper, and oaks provide more<br />

prey to owls during winter months than <strong>the</strong><br />

o<strong>the</strong>r three communities. However, temporal<br />

peaks of prey cycles are not correlated among<br />

<strong>the</strong>se communities. That is, when prey are<br />

abundant in mixed-conifer <strong>for</strong>est one year and<br />

low a subsequent year, an opposite pattern may<br />

occur in a different vegetation community. The<br />

asynchrony of abundance among prey species,<br />

vegetation communities, and time may provide a<br />

buffer against <strong>the</strong> effects of extreme oscillation in<br />

prey cycles. This implies maintaining a mixture<br />

of vegetation communities within <strong>the</strong> owl’s<br />

<strong>for</strong>aging range.<br />

Results of both studies also showed that <strong>the</strong><br />

owl’s food is most abundant during <strong>the</strong> summer<br />

when young are being raised. Decreases in prey<br />

biomass occur from late-fall through winter.<br />

Seasonal decreases, like <strong>the</strong>se, are typical of small<br />

mammal populations. Un<strong>for</strong>tunately, explanations<br />

<strong>for</strong> fluctuations in small mammal populations<br />

are equally variable and include artificial<br />

random patterns in population data, wea<strong>the</strong>r<br />

changes, behavioral mechanisms, predation, and<br />

age-related effects on demographic processes<br />

(e.g. fecundity, dispersal; see reviews by Conley<br />

and Nichols 1978, Finerty 1980).<br />

Although <strong>the</strong> reasons <strong>for</strong> seasonal declines in<br />

prey and <strong>the</strong>ir ramifications on <strong>the</strong> owl have not<br />

been quantified, conditions that increase winter<br />

food resources will likely improve conditions <strong>for</strong><br />

<strong>the</strong> owl. For example, Hirons (1985) has shown<br />

that large body reserves of fat and protein are<br />

essential during incubation by female tawny owls<br />

<strong>for</strong> successful reproduction. Abundant prey<br />

populations during winter and early spring<br />

periods increase <strong>the</strong> likelihood of egg laying and<br />

decrease <strong>the</strong> rate of nest abandonment (Hirons<br />

1985).<br />

Obviously, <strong>the</strong>re will be little recourse <strong>for</strong><br />

enhancing prey populations if factors like precipitation,<br />

temperature, or amount of snow


Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Figure Figure Figure 5.6. 5.6. Trends in average density (no./ha) of common prey of <strong>the</strong> <strong>Mexican</strong> spotted owl in (a)<br />

(a)<br />

mixed-conifer <strong>for</strong>est, (b) (b) montane meadows, and (c) (c) ponderosa pine-pinyon-juniper-oak woodlands,<br />

Sacramento Mountains, New Mexico. Summer (1992-93) values are based on 6 spatial replicates;<br />

summer (1991), fall, and winter values are based on 2 replicates.<br />

29


cover are <strong>the</strong> major causes of decline in prey<br />

populations. The degree to which habitat manipulation<br />

can ameliorate against wea<strong>the</strong>r effects<br />

or enhance prey availability requires investigation<br />

and is encouraged. Successful manipulation<br />

may provide a potent tool <strong>for</strong> recovering <strong>Mexican</strong><br />

spotted owls or o<strong>the</strong>r predators in <strong>the</strong> future.<br />

PREY PREY HABITAT<br />

HABITAT<br />

Ensuring adequate food <strong>for</strong> <strong>the</strong> owl requires<br />

conserving and possibly restoring habitat of <strong>the</strong><br />

owl’s prey. Ideally, to succeed, those features of<br />

<strong>the</strong> environment that consistently cause increases<br />

in abundance and availability of desired prey<br />

species should be identified. In reality, few<br />

habitat studies are so revealing or precise (see<br />

Verner et al. 1986). However, habitat studies<br />

often do identify patterns and descriptive correlates<br />

of animal distribution or abundance.<br />

Although crude, this type of in<strong>for</strong>mation frequently<br />

is all that is available <strong>for</strong> predicting <strong>the</strong><br />

outcome of planning decisions or management<br />

prescriptions.<br />

In <strong>the</strong> absence of cause-effect relationships<br />

among <strong>the</strong> owl’s prey and its habitat, we present<br />

habitat correlates <strong>for</strong> <strong>the</strong> distribution of several<br />

common prey species. This in<strong>for</strong>mation was<br />

determined from <strong>the</strong> same studies of prey<br />

abundance conducted in nor<strong>the</strong>rn Arizona<br />

(Block and Ganey, unpublished data) and <strong>the</strong><br />

Sacramento Mountains (Ward et al., unpublished<br />

data).<br />

Nor<strong>the</strong>rn Nor<strong>the</strong>rn Nor<strong>the</strong>rn Arizona Arizona Arizona - - - Pine-oak Pine-oak Pine-oak Forest Forest<br />

Forest<br />

Field Field Methods Methods and and Statistical Statistical Analyses Analyses<br />

Analyses<br />

Habitat-sampling plots were established as a<br />

5-m [16.5-ft) radius centered at each trapping<br />

station (n = 1,260). Cover by grass, <strong>for</strong>bs, rock,<br />

dead woody debris of three size classes (10 cm [3.9<br />

in]), and live woody vegetation at four height<br />

strata (2-5 m<br />

[6.6-16.4 ft], >5 m [16.4 ft]) were estimated as<br />

<strong>the</strong> percentage of 10 point intercepts (at 1-m<br />

[3.3 ft] intervals along a randomly oriented<br />

transect) covered by each of <strong>the</strong>se variables. Tree<br />

Volume II/Chapter 5<br />

30<br />

diameters were measured with a dbh tape;<br />

heights were measured with a clinometer. Shrub<br />

and slash pile heights were measured with a<br />

meter stick. Mid-point diameters and lengths of<br />

logs within <strong>the</strong> plot were measured with a<br />

measuring tape. Numbers of trees and shrubs by<br />

species were recorded. Slope was measured with<br />

a clinometer and aspect with a compass.<br />

Deer mice, brush mice, and <strong>Mexican</strong><br />

woodrats were captured in sufficient numbers to<br />

permit habitat analyses. Stations where each<br />

species was captured were contrasted with those<br />

where it was not captured using analysis of<br />

variance with owl territory as a blocking factor.<br />

Two o<strong>the</strong>r analyses were also conducted. Logistic<br />

regression was used to examine differences<br />

between stations where each species was and was<br />

not captured that may not have been identified<br />

in <strong>the</strong> ANOVA. Stepwise multiple linear regression<br />

(Draper and Smith 1981) was used to<br />

evaluate relationships between small mammal<br />

population levels and habitat characteristics. The<br />

dependent variable was <strong>the</strong> number of trapping<br />

stations/grid where <strong>the</strong> species was captured.<br />

Independent variables were habitat characteristics<br />

aggregated across <strong>the</strong> grid.<br />

Results<br />

Results<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Deer mice used more open sites, on gentler<br />

slopes, and with less shrub and midstory canopy,<br />

smaller densities of Gambel oak trees and shrubs,<br />

but more slash piles and greater litter depth than<br />

stations where it was not captured (Table 5.11).<br />

In contrast, brush mice and <strong>Mexican</strong> woodrats<br />

both used sites characterized by greater slopes,<br />

low vegetation cover, sparser tree canopy (>5 m<br />

[16.4 ft]) cover, more Gambel oak shrubs, and<br />

greater log volume than areas where <strong>the</strong>y were<br />

not captured (Table 5.11). Fur<strong>the</strong>r, brush mice<br />

used areas with greater Gambel oak tree density<br />

and basal area and less ponderosa pine basal area<br />

than found at stations where it was not captured.<br />

Tree basal area at sites where <strong>Mexican</strong> woodrats<br />

were captured was significantly less and total<br />

rock cover was significantly greater than at sites<br />

where it was not captured.<br />

Generally, results from logistic regression<br />

analyses corroborated results of <strong>the</strong> univariate<br />

analyses. Deer mice were associated with areas of


31<br />

Table Table 5.11. 5.11. Descriptive statistics (means with SE in paren<strong>the</strong>ses) of selected habitat variables characterizing habitats of common mammalian prey of<br />

<strong>Mexican</strong> spotted owls (4 territories) in ponderosa pine-Gambel oak <strong>for</strong>ests, nor<strong>the</strong>rn Arizona, 1990-1992.


32<br />

Table Table 5.11. 5.11. (continued)<br />

FPO


little slope, Gambel oak, and mid-story (2-5 m<br />

[6.6-16.4 ft]) vegetation. Brush mice and <strong>Mexican</strong><br />

woodrats used areas with greater shrub<br />

density, rock cover, log volume, and Gambel oak<br />

cover.<br />

In <strong>the</strong> multiple regression analysis, no<br />

variable entered <strong>the</strong> model <strong>for</strong> deer mice, suggesting<br />

that it is a habitat generalist. Two variables,<br />

shrub density and grass height, entered <strong>the</strong><br />

regression model <strong>for</strong> brush mice. These two<br />

variables alone explained >96% of <strong>the</strong> variation<br />

in <strong>the</strong> data set. For <strong>Mexican</strong> woodrats, shrub<br />

density and slope entered <strong>the</strong> model and explained<br />

>92% of <strong>the</strong> variation in numbers. Thus,<br />

both brush mice and <strong>Mexican</strong> woodrats were<br />

apparently more stenotypic in habitat than deer<br />

mice and closely associated with shrub cover<br />

provided by Gambel oak and New <strong>Mexican</strong><br />

locust.<br />

Sacramento Sacramento Mountains<br />

Mountains<br />

Field Field Field Methods Methods and and Statistical Statistical Analyses Analyses<br />

Analyses<br />

Habitat characteristics were measured within<br />

circular plots of 5-m [16.4-ft] radius centered at<br />

about 80% of <strong>the</strong> trap stations (n = 1,416).<br />

Stations were stratified by vegetation community:<br />

mesic <strong>for</strong>est (n = 583), xeric <strong>for</strong>est<br />

(n = 578), and meadow (n = 255). Within each<br />

plot, methods similar to those described above<br />

were used with <strong>the</strong> following exception: tree<br />

basal area and density were estimated with a<br />

plotless method using 10- and 20-factor prisms<br />

ra<strong>the</strong>r than <strong>the</strong> plot methods used in <strong>the</strong> ponderosa<br />

pine-Gambel oak <strong>for</strong>est of nor<strong>the</strong>rn Arizona.<br />

Also, cover of woody vegetation by height<br />

strata was not recorded in <strong>the</strong> Sacramento<br />

Mountains.<br />

Deer mice, brush mice, <strong>Mexican</strong> woodrats,<br />

long-tailed voles, and <strong>Mexican</strong> voles were captured<br />

in sufficient numbers to permit habitat<br />

analyses. Statistical analyses <strong>for</strong> each species were<br />

done separately by vegetation community.<br />

Habitat characteristics of trap stations where a<br />

species was captured were contrasted with<br />

stations where <strong>the</strong>y were not captured using<br />

Student’s t-test. A random sample of unused<br />

stations equal to <strong>the</strong> number of used stations was<br />

Volume II/Chapter 5<br />

33<br />

selected to meet <strong>the</strong> assumption of equal sample<br />

size except <strong>for</strong> <strong>Mexican</strong> voles in meadows, which<br />

were captured at >75% of <strong>the</strong> trap stations.<br />

Consequently, a subset of used stations equal to<br />

<strong>the</strong> available number of unused stations was<br />

randomly selected (n = 62) <strong>for</strong> conducting <strong>the</strong><br />

analysis. Logistic regression was also used to<br />

determine <strong>the</strong> relative value of variables in<br />

distinguishing between stations where <strong>the</strong><br />

animal was and was not captured. This second<br />

analysis was used to detect o<strong>the</strong>r variables that<br />

might identify habitat correlates not identified<br />

by use of Student t-tests.<br />

Results<br />

Results<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Mesic esic F FFor<br />

F or orests. or ests. ests.—Microhabitat ests.<br />

analyses were<br />

possible <strong>for</strong> deer mice, <strong>Mexican</strong> woodrats, and<br />

long-tailed and <strong>Mexican</strong> voles. Deer mice were<br />

captured in areas with little herbaceous cover and<br />

extensive exposed soil (Table 5.12). Results from<br />

<strong>the</strong> logistic regression indicated that deer mice<br />

used areas with less herbaceous cover and greater<br />

densities of live conifer trees. <strong>Mexican</strong> woodrats<br />

used areas that had greater shrub but less herbaceous<br />

cover (Table 5.13). Shrub cover was <strong>the</strong><br />

only variable included in <strong>the</strong> stepwise logistic<br />

regression model of habitat use by <strong>Mexican</strong><br />

woodrats. Long-tailed voles occurred in areas<br />

characterized by less slope, less tree cover and<br />

exposed soil but greater herbaceous cover, fewer<br />

stumps, greater shrub numbers, and fewer<br />

conifer snags (Table 5.14). According to <strong>the</strong><br />

logistic regression model, long-tailed voles used<br />

areas with less exposed soil, greater shrub density,<br />

and less slope. <strong>Mexican</strong> voles used sites with less<br />

shrub, tree and exposed ground cover, greater<br />

herbaceous cover, fewer shrubs, fewer conifer<br />

seedlings and saplings, lower density and less<br />

basal area of deciduous trees (Table 5.15).<br />

<strong>Mexican</strong> voles used sites with less tree cover and<br />

lower deciduous tree density according to <strong>the</strong><br />

logistic regression model.<br />

Xeric eric eric FF<br />

For FF<br />

or or orests. or ests. ests.—Habitat ests. analyses were<br />

possible <strong>for</strong> deer mice, brush mice, <strong>Mexican</strong><br />

woodrats, and <strong>Mexican</strong> voles. As in mesic <strong>for</strong>ests,<br />

deer mice occupying xeric <strong>for</strong>ests were captured<br />

in areas with more exposed soil than at<br />

noncapture stations (Table 5.12). In contrast<br />

with mesic <strong>for</strong>ests, deer mice in xeric <strong>for</strong>ests were


34<br />

Table Table 5.12. 5.12. Habitat characteristics used by deer mice in three vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico. Only those variables<br />

that differed significantly between trap stations where deer mice were and were not captured are reported.<br />

FPO


35<br />

Table Table 5.13. 5.13. Habitat characteristics used by <strong>Mexican</strong> woodrats in two vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico. Only those<br />

variables that differed significantly between trap stations where woodrats were and were not captured are reported.<br />

FPO


36<br />

Table Table 5.14. 5.14. Habitat characteristics used by long-tailed voles in two vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico. Only those<br />

variables that differed significantly between trap stations where long-tailed voles were and were not captured are reported.<br />

FPO


37<br />

Table Table 5.15. 5.15. Habitat characteristics used by <strong>Mexican</strong> voles in three vegetation communities of <strong>the</strong> Sacramento Mountains, New Mexico. Only those<br />

variables that differed significantly between trap stations where <strong>Mexican</strong> voles were and were not captured are reported.<br />

FPO


found on flatter areas with less rock but greater<br />

shrub cover, and greater density and basal area of<br />

live conifer trees (t-test, P < 0.05). Logistic<br />

regression identified slope, density of live conifer<br />

trees, litter depth, and number of shrub species<br />

as useful variables <strong>for</strong> distinguishing deer mouse<br />

habitat. Brush mice were found in areas having<br />

less tree cover, greater rock cover, shallower litter<br />

depth, greater densities of shrubs, gray oak, and<br />

conifer seedlings and saplings, but fewer conifer<br />

trees (Table 5.16). Logistic regression results<br />

indicated that brush mice used areas with shallower<br />

litter depth, greater shrub density, and<br />

fewer conifer trees. <strong>Mexican</strong> woodrats were<br />

captured in areas with greater shrub cover and<br />

density, and greater cover by gray oak (Table<br />

5.13). Shrub cover was <strong>the</strong> only variable that<br />

discriminated capture from noncapture sites by<br />

logistic regression. <strong>Mexican</strong> voles were found at<br />

sites with greater herbaceous cover and height,<br />

and greater density and basal area of conifer<br />

snags (Table 5.15). Logistic regression identified<br />

two variables, density of conifer snags and height<br />

of herbaceous cover, as habitat discriminates <strong>for</strong><br />

<strong>Mexican</strong> voles occupying xeric <strong>for</strong>ests.<br />

Meado eado eadows. eado ws. ws.—Deer ws. mice, long-tailed voles,<br />

and <strong>Mexican</strong> voles were captured frequently<br />

enough to permit analyses. As in both <strong>for</strong>est<br />

types, deer mice occupying meadows were<br />

captured in areas with more exposed soil (Table<br />

5.12). These mice also used areas with steeper<br />

slopes, shallower litter depth, and more shrub<br />

species (Table 5.12). Logistic regression indicated<br />

that only steeper slopes could be used to<br />

distinguish between capture and noncapture sites<br />

of deer mice. Long-tailed voles were captured at<br />

stations with less herbaceous cover, greater shrub<br />

density, and a greater number of shrub species<br />

(Table 5.14). Even though cover by herbaceous<br />

vegetation was less at capture stations, it still<br />

averaged 90.0% (SE = 2.8). Shrub density and<br />

herbaceous cover were <strong>the</strong> two variables that<br />

separated capture from noncapture sites using<br />

logistic regression. <strong>Mexican</strong> voles were found on<br />

flatter areas with greater herbaceous cover and<br />

height, greater litter depth, and less exposed<br />

ground (Table 5.15). Logistic regression identified<br />

slope, height of herbaceous vegetation, and<br />

litter depth as habitat variables associated with<br />

<strong>the</strong> presence of <strong>Mexican</strong> voles in meadows.<br />

Volume II/Chapter 5<br />

38<br />

DISTURBANCE DISTURBANCE EFFECTS EFFECTS ON<br />

ON<br />

OWL OWL PREY PREY AND AND HABITAT<br />

HABITAT<br />

The distribution and abundance of <strong>the</strong><br />

spotted owl’s prey are influenced by both natural<br />

and anthropogenic factors. Though all factors to<br />

some degree have <strong>for</strong>mative character, many are<br />

more accurately described as disturbance factors.<br />

Definitions of disturbance and <strong>the</strong> magnitude of<br />

associated effects vary according to ecological<br />

scale and conditions. Here, we briefly discuss<br />

fire, tree harvesting, and livestock grazing because<br />

<strong>the</strong>se activities operate at spatial scales<br />

likely to influence spotted owls. Specifically, <strong>the</strong>y<br />

may determine whe<strong>the</strong>r an owl occupies and<br />

reproduces in a given area. Important to this<br />

discussion is <strong>the</strong> element of human control over<br />

<strong>the</strong>se disturbance activities.<br />

We know little about direct cause-effect<br />

relationships of most natural and anthropogenic<br />

disturbance factors on owl prey populations.<br />

Correlative in<strong>for</strong>mation allows us to infer some<br />

effects but most published research is from areas<br />

outside <strong>the</strong> range of <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Their applicability to southwestern conditions is<br />

uncertain. Fur<strong>the</strong>rmore, effects of disturbance<br />

will vary by species, time, and space. Consequently,<br />

it is important to view disturbance at<br />

each of <strong>the</strong>se scales. The issue becomes even<br />

more complicated when one considers <strong>the</strong><br />

synergistic and cumulative effects of multiple<br />

disturbances. The latter scenario is more likely<br />

<strong>the</strong> norm than <strong>the</strong> exception.<br />

Fire Fire<br />

Fire<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Generalizing about <strong>the</strong> effects of fire on <strong>the</strong><br />

owl’s prey is impossible <strong>for</strong> a number of reasons<br />

including variations in fire characteristics and in<br />

prey habitat. Fire intensity, size, and behavior are<br />

influenced by numerous factors such as vegetation<br />

type, moisture, fuel loads, wea<strong>the</strong>r, season,<br />

and topography. Data presented in <strong>the</strong> previous<br />

sections illustrate how macrohabitat and microhabitat<br />

associations of <strong>the</strong> owl’s prey vary both<br />

geographically within a species and among<br />

species.<br />

Fire can effectively alter vegetation structure<br />

and composition <strong>the</strong>reby affecting small mam-


39<br />

Table Table 5.16. 5.16. Habitat characteristics used by brush mice in xeric <strong>for</strong>ests of <strong>the</strong> Sacramento Mountains, New Mexico. Only those variables that differed<br />

significantly between trap stations where brush mice were and were not captured are reported.<br />

FPO


mal habitats. Population responses by small<br />

mammals to fire-induced changes in <strong>the</strong>ir<br />

habitat vary. For example, deer mouse populations<br />

might increase immediately following fire<br />

and <strong>the</strong>n decrease through time (Peterson et al.<br />

1985, Kaufman et al. 1988). Based on limited<br />

sampling ef<strong>for</strong>ts restricted to one location,<br />

Campbell et al. (1977) noted that populations of<br />

peromyscid mice decreased immediately following<br />

fire in an Arizona ponderosa pine <strong>for</strong>est that<br />

removed one-fourth (moderately burned) to<br />

two-thirds (severely burned) of <strong>the</strong> basal area;<br />

populations <strong>the</strong>n returned to prefire numbers<br />

two years following <strong>the</strong> burn. Fur<strong>the</strong>r, no differences<br />

were found in rodent populations between<br />

moderately and severely burned areas. They<br />

concluded that <strong>the</strong> effects of <strong>the</strong> fire that <strong>the</strong>y<br />

studied were short-term, and <strong>the</strong> short-term<br />

positive numerical responses by mice were<br />

attributed to an increase in <strong>for</strong>age, particularly<br />

grasses and <strong>for</strong>bs.<br />

However, we suspect that effects of more<br />

intense stand-replacing fires that dramatically<br />

alter <strong>for</strong>est structure and move <strong>the</strong> system to<br />

earlier seral stages would have longer-term effects<br />

on rodent populations. Likely, early successional<br />

species (such as <strong>the</strong> deer mouse) and those that<br />

require open habitats with a well developed<br />

herbaceous understory (such as microtine voles<br />

and pocket gophers) would benefit. In contrast,<br />

species that require a wooded or <strong>for</strong>ested overstory<br />

would exhibit population declines. The net<br />

effect of such fires on spotted owls is unclear. A<br />

fire that removes <strong>the</strong> tree canopy would likely<br />

render <strong>the</strong> area unusable <strong>for</strong> <strong>for</strong>aging because of<br />

how spotted owls <strong>for</strong>age. But if <strong>the</strong> spatial extent<br />

of crown loss is limited, a mosaic is created that<br />

could provide a diversity of prey <strong>for</strong> <strong>the</strong> owl and<br />

actually be beneficial.<br />

Clearly, research is required to determine <strong>the</strong><br />

effects of fire on spotted owl prey. Because owl<br />

prey species evolved in ecosystems where fire was<br />

a natural process, we assume that <strong>the</strong>se species<br />

survive and some even benefit from <strong>the</strong> occurrence<br />

of fire. Fire has been excluded from most<br />

southwestern ecosystems during <strong>the</strong> 20th century,<br />

resulting in systems where fire behavior<br />

may deviate substantially from natural conditions.<br />

Effects of fire on small mammals under<br />

present environmental conditions are unclear.<br />

Volume II/Chapter 5<br />

40<br />

Timber Timber Harvest Harvest and<br />

and<br />

Fuelwood Fuelwood Removal Removal<br />

Removal<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Numerous silvicultural methods are used to<br />

harvest timber in <strong>the</strong> Southwest. These include<br />

both even-aged (predominantly shelterwood<br />

systems) and uneven-aged (e.g., single tree,<br />

group selection systems) management. Fur<strong>the</strong>r,<br />

<strong>the</strong>se systems are applied differently according to<br />

site characteristics and management objectives.<br />

Given <strong>the</strong> various scenarios <strong>for</strong> silviculture,<br />

generalizing about <strong>the</strong>ir effects on prey populations<br />

is not possible.<br />

Tree removal, whe<strong>the</strong>r to harvest sawlogs or<br />

fuelwood, will affect natural ecosystem processes<br />

in numerous obvious and obscure ways. Certainly,<br />

removal of mast-producing trees (e.g.,<br />

pinyon pine, juniper, oak) reduces food availability<br />

<strong>for</strong> several of <strong>the</strong> owl’s prey species. Also,<br />

removal of tree biomass from <strong>the</strong> site will interrupt<br />

both nutrient cycling and energy flow. The<br />

effects of altering <strong>the</strong>se processes on owl prey<br />

populations is unknown, but <strong>the</strong> disturbance<br />

may likely benefit some species while negatively<br />

affecting o<strong>the</strong>rs. Tree removal and accompanying<br />

site disturbance during and following removal,<br />

plus residual disturbance such as soil compaction,<br />

increased erosion, and creation of slash<br />

piles, will directly alter habitats of many prey<br />

species. Block and Ganey (unpublished data)<br />

found more deer mice in areas with slash than<br />

areas without slash. In <strong>the</strong> same general area,<br />

Goodwin and Hunger<strong>for</strong>d (1979) noted that<br />

brush mice and <strong>Mexican</strong> woodrats used long<br />

windrows of slash following logging. Block and<br />

Ganey (unpublished data) sampled <strong>the</strong> same<br />

areas 18-20 years after <strong>the</strong> Goodwin and<br />

Hunger<strong>for</strong>d study and captured few woodrats<br />

and brush mice. This suggests a temporary<br />

benefit of slash piles in that <strong>the</strong>y may provide a<br />

short-term habitat component that is not used as<br />

<strong>the</strong> slash becomes compressed and decomposes.<br />

As noted above, effects of tree removal on<br />

small mammals varies by prescription and site<br />

characteristics. The effects are somewhat scale<br />

dependent. Prescriptions <strong>for</strong> even-aged,<br />

shelterwood cuts or clearcuts effectively return<br />

large areas, <strong>the</strong> size of stands or greater, to earlier<br />

seral stages. Small mammal community structure


is correlated to plant seral stage (Fox 1990,<br />

Kirkland 1990). Thus, following even-aged<br />

management, populations of some species will<br />

respond positively and o<strong>the</strong>rs negatively. As time<br />

progresses and vegetation proceeds through<br />

succession, population dynamics of species will<br />

change in response to changing habitat conditions<br />

and community structure. Areas subjected<br />

to even-aged management may not provide<br />

appropriate <strong>for</strong>aging habitat <strong>for</strong> <strong>the</strong> spotted owl<br />

(see Ganey and Dick 1995). This means changes<br />

to small mammal populations within <strong>the</strong> harvested<br />

areas may not affect <strong>the</strong> owl prey base<br />

directly because those animals may not be<br />

available to <strong>the</strong> owl. However, <strong>the</strong> same small<br />

mammal populations may provide a source of<br />

individuals that disperse into adjacent owl<br />

<strong>for</strong>aging habitat.<br />

Uneven-aged management would likely be<br />

used over large areas and does not create small<br />

stands, but ra<strong>the</strong>r it creates groups or clumps.<br />

Mosaics of habitat provide diverse plant communities<br />

and o<strong>the</strong>r conditions that, collectively, can<br />

support a rich diversity of fauna. Populations of<br />

different species may respond variably to aspects<br />

of <strong>the</strong> mosaic pattern, such as conditions within<br />

each patch type, patch size and shape, and<br />

interspersion and juxtaposition of <strong>the</strong>se patches.<br />

Mosaic patterns resulting from timber management<br />

prescriptions such as single-tree or group<br />

selection cuts may in some ways mimic natural<br />

disturbance patterns and create canopy gaps.<br />

This does not imply that effects of silviculture<br />

are equivalent to those of natural disturbance,<br />

only that <strong>the</strong> resultant spatial patterns are somewhat<br />

similar.<br />

Clearly, research is needed to determine<br />

cause-effect relationships of tree removal on<br />

spotted owl prey populations, <strong>the</strong> hunting ability<br />

of <strong>the</strong> owl, and <strong>the</strong> mosaic patterns which best<br />

conserve owl populations. Such research will<br />

entail experiments conducted at varying spatial<br />

and temporal scales to understand <strong>the</strong> true<br />

magnitude of <strong>the</strong> effects. Until <strong>the</strong>se experiments<br />

are conducted, effects of tree removal on<br />

prey habitat and populations must be based on<br />

speculation and conjecture. Future management<br />

conducted within a scientific and experimental<br />

context may provide a means <strong>for</strong> establishing<br />

Volume II/Chapter 5<br />

41<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

cause-effect relationships (see USDI 1995:<br />

Activity-Specific Research Section, Part III.E.).<br />

Grazing<br />

Grazing<br />

The effects of livestock and wildlife grazing<br />

on spotted owl prey populations and <strong>the</strong>ir<br />

habitats is also a complex issue. Impacts can vary<br />

according to grazing species; degree of use,<br />

including numbers of grazers, grazing intensity,<br />

grazing frequency, and timing of grazing; habitat<br />

type and structure; and plant or prey species<br />

composition. It is well documented and intuitive<br />

that repetitive, excessive grazing of plant communities<br />

by livestock can significantly alter plant<br />

species density, composition, vigor, regeneration,<br />

above or below ground phytomass, soil properties,<br />

nutrient flow, water quality, and ultimately<br />

lead to desertification when uncontrolled<br />

(Kauffman and Krueger 1984, Orodho et al.<br />

1990, Vallentine 1990, Milchunas and<br />

Lauenroth 1993). These effects can have both<br />

direct and indirect adverse impacts on animal<br />

species that are dependent on plants <strong>for</strong> food<br />

and cover. However, moderate to light grazing<br />

can benefit some plant and animal species under<br />

certain conditions and in certain environments,<br />

maintain communities in certain seral stages,<br />

and increase primary productivity (Reynolds<br />

1980, Hanley and Page 1982, Kauffman and<br />

Krueger 1984, McNaughton 1993). Fur<strong>the</strong>r,<br />

direct influences of livestock on plant communities<br />

are not always reflected in small mammal<br />

communities (Grant et al. 1982). Thus, any<br />

generalizations presented here should not be<br />

construed as absolute; <strong>the</strong>re are exceptions.<br />

No studies document <strong>the</strong> direct and indirect<br />

effects of livestock and wildlife grazing on <strong>the</strong><br />

<strong>Mexican</strong> spotted owl or its prey (see reviews by<br />

USDA Forest Service 1994, Utah <strong>Mexican</strong><br />

<strong>Spotted</strong> <strong>Owl</strong> Technical Team 1994). We found<br />

only one study that specifically investigated <strong>the</strong><br />

effects of livestock grazing on an upland <strong>for</strong>est<br />

owl, <strong>the</strong> tawny owl (Putnam 1986). However,<br />

<strong>the</strong> design and limited sampling ef<strong>for</strong>t of this<br />

study prevents extrapolation to <strong>the</strong> <strong>Mexican</strong><br />

spotted owl. Interpretive extrapolations are<br />

fur<strong>the</strong>r hampered because most livestock-effects<br />

studies on small mammals have been conducted


in shrub-steppe, grassland, or riparian communities<br />

of arid or semi-arid regions (Kauffman and<br />

Krueger 1984, Milchunas and Lauenroth 1993).<br />

Except <strong>for</strong> <strong>the</strong> possible use of <strong>the</strong>se areas <strong>for</strong><br />

dispersing or wintering, such areas are not<br />

typically used by <strong>Mexican</strong> spotted owls.<br />

Despite <strong>the</strong> dearth of specific in<strong>for</strong>mation<br />

about spotted owls and grazing, <strong>the</strong>re exists some<br />

knowledge regarding <strong>the</strong> effects of livestock<br />

grazing on small mammals frequently consumed<br />

by spotted owls and regarding mesic or montane<br />

plant communities inhabited by <strong>the</strong> owl’s prey.<br />

Relevant studies include Hanley and Page<br />

(1982), Medin and Clary (1990), Schultz and<br />

Leininger (1991), and Szaro (1991), and are<br />

briefly summarized below.<br />

In a 250,000-ha (619,250-ac) area of <strong>the</strong><br />

Great Basin, Hanley and Page (1982) found that<br />

livestock grazing (primarily cattle with an<br />

average of 0.16 animal unit month, April to<br />

October during each of four consecutive years)<br />

generally increased shrub composition and<br />

decreased perennial herbs and grasses (primarily<br />

tall bunchgrasses). However, <strong>the</strong>se effects depended<br />

upon community type. Shrubs increased<br />

in grazed wet meadows and willow-riparian<br />

communities, whereas tree-<strong>for</strong>m willows decreased<br />

by 60%. Herbaceous layers were 30-50<br />

cm (11.8-19.7 in) deep in ungrazed meadows<br />

compared to <strong>the</strong> “mowed” appearance of grazed<br />

meadows. Reduction of graminoids and <strong>for</strong>bs<br />

increased structural diversity in mesic habits by<br />

increasing shrubs but decreased structural<br />

diversity in shrub-dominated, xeric habitats. The<br />

authors postulated that <strong>the</strong> loss of perennial<br />

grasses and herbs caused lower numbers of desert<br />

woodrats, and long-tailed and mountain voles in<br />

grazed meadows by decreasing food sources and<br />

cover. The authors also reported a greater number<br />

of Great Basin pocket mice, deer mice, and<br />

least chipmunks in mesic habitats grazed by<br />

livestock compared to ungrazed, mesic habitats.<br />

However, <strong>the</strong>se same three species decreased in<br />

xeric habitats grazed by livestock.<br />

Schultz and Leininger (1991) examined <strong>the</strong><br />

effects of cattle exclusion along one 5-km (3.1mi)<br />

length of a riparian community in<br />

northcentral Colorado. During <strong>the</strong> summers of<br />

1985 and 1986, plant and small mammal<br />

communities were sampled <strong>for</strong> comparison<br />

Volume II/Chapter 5<br />

42<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

between adjacent grazed areas and three<br />

exclosures established in 1956 and 1959. <strong>Plan</strong>t<br />

communities varied from grass-herb-sedge<br />

lowlands to ponderosa pine uplands. Observed<br />

differences included greater vascular vegetation<br />

cover, graminoid cover, and shrub cover in <strong>the</strong><br />

exclosures. Litter accumulation in <strong>the</strong> exclosures<br />

was nearly twice that of grazed areas, and willow<br />

canopy cover was 8.5 times greater (Schultz<br />

1990). Deer mice were significantly more<br />

abundant in grazed areas and western jumping<br />

mice were significantly more abundant in<br />

ungrazed exclosures. Seven species of small<br />

mammals were found in grazed and exclosed<br />

areas, although species composition in each area<br />

was slightly different. Deer mice, least chipmunks,<br />

masked shrews, dusky shrews, and<br />

western jumping mice were found in both areas.<br />

Long-tailed and mountain voles were not observed<br />

in grazed areas. Nor<strong>the</strong>rn pocket gophers<br />

and golden-mantled ground squirrels were not<br />

found in exclosures.<br />

Medin and Clary (1990) compared small<br />

mammal populations between a riparian habitat<br />

seasonally grazed by cattle to an adjacent 122-ha<br />

(302.2-ac) riparian exclosure which had been<br />

protected from livestock grazing <strong>for</strong> <strong>the</strong> previous<br />

14 years. They found that mountain voles were<br />

four times more abundant within <strong>the</strong> ungrazed<br />

exclosure as compared to <strong>the</strong> grazed riparian<br />

habitat. In addition, vagrant shrews, water<br />

shrews, and nor<strong>the</strong>rn pocket gophers were only<br />

trapped within <strong>the</strong> ungrazed habitat. Consequently,<br />

<strong>the</strong>y noted that small mammal species<br />

richness and species diversity were higher within<br />

<strong>the</strong> ungrazed exclosure despite <strong>the</strong> fact that total<br />

population density of small mammals was a<br />

third higher in <strong>the</strong> grazed portion due to <strong>the</strong><br />

high numbers of deer mice.<br />

Szaro (1991) examined <strong>the</strong> effects of grazing<br />

along <strong>the</strong> Rio de las Vacas, a montane stream at<br />

8,528 ft (2,600 m) elevation in <strong>the</strong> San Pedro<br />

Parks Wilderness Area, New Mexico. Terrestrial<br />

fauna within two 900 x 50 m (2,952 x 164 ft)<br />

livestock exclosures were sampled and compared<br />

to downstream private lands that were continuously<br />

grazed. Small mammals were sampled each<br />

month, June through September, in 1985 and<br />

1986. He found greater amounts of herbaceous<br />

vegetation in <strong>the</strong> exclosures. Greater numbers of


small mammal captures were observed in both<br />

exclosures in 1985 compared to grazed areas<br />

downstream. Small mammal captures were<br />

reduced and indistinguishable in 1986 but some<br />

cattle trespass had occurred into <strong>the</strong> control<br />

exclosures. Fewer species were captured in <strong>the</strong><br />

grazed areas. Mountain voles and deer mice were<br />

found in both grazed and ungrazed situations,<br />

whereas least chipmunks and golden-mantled<br />

ground squirrels were found only in ungrazed<br />

areas.<br />

O<strong>the</strong>r studies have shown similar results:<br />

lack of a numerical decrease by deer mice to<br />

grazing (Reynolds 1980), and significant decrease<br />

in voles caused by grazing induced loss of<br />

cover in mesic habitats (Grant et al. 1982). If<br />

<strong>the</strong>se general patterns can be applied to upland<br />

habitats of <strong>the</strong> Southwest, we would expect<br />

moderate to heavy grazing to decrease populations<br />

of voles and improve conditions <strong>for</strong> deer<br />

mice in meadow habitats. Such decreases could<br />

negatively influence spotted owls occupying<br />

areas in <strong>the</strong> Upper Gila Mountains, Basin and<br />

Range - East, and portions of o<strong>the</strong>r RUs where<br />

voles are common prey or used as alternative<br />

food sources when o<strong>the</strong>r prey species are diminished.<br />

Increases in deer mouse abundance in<br />

meadows probably would not offset decreases in<br />

vole numbers because voles provide greater<br />

biomass per individual and per unit of area. Loss<br />

of perennial grass cover in xeric communities<br />

used by owls, specifically, ponderosa pine <strong>for</strong>est<br />

and pinyon-juniper woodlands, due to grazing<br />

may reduce deer mice and <strong>Mexican</strong> vole populations.<br />

The reduction of <strong>the</strong>se prey species in<br />

xeric communities could be more critical than<br />

that in meadows if xeric habitats are necessary<br />

<strong>for</strong> winter <strong>for</strong>aging by <strong>the</strong> owls.<br />

Finally, high intensity grazing in riparian<br />

communities during <strong>the</strong> fall and winter seasons<br />

where grass seedheads may be totally removed<br />

can cause significant short-term decreases in<br />

small mammal populations (Kauffman et al.<br />

1983). Continued heavy grazing in upland or<br />

lowland riparian communities could <strong>the</strong>re<strong>for</strong>e<br />

greatly reduce <strong>the</strong> potential <strong>for</strong> utilization by<br />

<strong>for</strong>aging, dispersing or wintering spotted owls.<br />

Volume II/Chapter 5<br />

43<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

CONCLUSIONS<br />

CONCLUSIONS<br />

<strong>Mexican</strong> spotted owls consume a variety of<br />

prey throughout <strong>the</strong>ir range but commonly eat<br />

small and medium-sized rodents. However, <strong>the</strong><br />

owl’s food habits vary according to geographic<br />

location. For example, spotted owls dwelling in<br />

canyons of <strong>the</strong> Colorado Plateau take more<br />

woodrats, and fewer voles and birds than do<br />

spotted owls from o<strong>the</strong>r areas. In contrast,<br />

spotted owls occupying mountain ranges with<br />

<strong>for</strong>est-meadow interfaces, such as <strong>the</strong> Basin and<br />

Range - East, Sou<strong>the</strong>rn Rocky Mountains -<br />

Colorado, and Upper Gila Mountains RUs, take<br />

more microtine voles The differences in diet<br />

likely reflect geographic variation in population<br />

densities and habitats of both <strong>the</strong> prey and <strong>the</strong><br />

owl.<br />

No strong rangewide relationships appeared<br />

in our analyses of <strong>the</strong> owl’s diet and reproduction.<br />

The relationship was positive and nearly<br />

significant when comparing <strong>the</strong> prevalence of<br />

<strong>the</strong> three most common prey (peromyscid mice,<br />

woodrats, and voles) in <strong>the</strong> diet and owl reproduction,<br />

implying that multiple species influence<br />

<strong>the</strong> owl's fitness. However, this generalization<br />

may not apply to owls in <strong>the</strong> Sacramento Mountains<br />

where <strong>the</strong> owl's reproduction appears most<br />

influenced by deer mouse abundance. In addition,<br />

<strong>the</strong> predominance of woodrats, both in diet<br />

frequency and biomass throughout much of <strong>the</strong><br />

owl’s range, suggests that a single prey may<br />

influence <strong>the</strong> owl’s fitness. O<strong>the</strong>r studies have<br />

shown positive associations between larger prey<br />

(e.g. woodrats) in <strong>the</strong> diet of nor<strong>the</strong>rn and<br />

Cali<strong>for</strong>nia spotted owls and its reproductive<br />

success (Barrows 1987, Thrailkill and Bias<br />

1989). The lack of more specific results should<br />

not imply that a simple relationship between<br />

diet and reproduction or o<strong>the</strong>r fitness measures<br />

does not exist. Ra<strong>the</strong>r, those relationships are<br />

more complex than are evident by <strong>the</strong> available<br />

in<strong>for</strong>mation and analytical approaches used in<br />

producing this report. In most cases, total prey<br />

biomass is likely more influential on <strong>the</strong> owl’s<br />

fitness than <strong>the</strong> abundance of any particular prey<br />

species. O<strong>the</strong>r factors worth exploring would be<br />

lag effects such as <strong>the</strong> effects of winter diet on<br />

breeding potential, prey diversity, or synergistic


effects of diet with factors like owl density,<br />

wea<strong>the</strong>r, and habitat.<br />

Studies conducted in four vegetation communities<br />

demonstrate that abundance of <strong>the</strong><br />

owl’s food varies according to habitat and time.<br />

In <strong>the</strong> Sacramento Mountains, <strong>the</strong> greatest prey<br />

biomass is found in high-elevation meadows<br />

occurring along riparian corridors. Common<br />

prey species occupying <strong>the</strong>se meadows are longtailed<br />

voles, <strong>Mexican</strong> voles, and deer mice.<br />

However, abundance alone does not necessarily<br />

connote availability. Availability infers cooccurrence<br />

of owl and prey plus coincidental<br />

vulnerability to predation and ability to capture.<br />

Successful capture of meadow-dwelling rodents<br />

may be restricted to areas near <strong>for</strong>est edges<br />

coinciding with <strong>the</strong> presence of <strong>for</strong>aging perches.<br />

Ra<strong>the</strong>r, meadow habitats may play an indirect<br />

role by producing high densities of prey that<br />

become available to owls following dispersal into<br />

adjacent <strong>for</strong>ests. <strong>Owl</strong>s in <strong>the</strong> Sacramento Mountains<br />

consume a moderate-to-large proportion of<br />

voles during years of high vole density.<br />

Summer prey biomass in mesic (mixedconifer)<br />

<strong>for</strong>ests can be greater than in xeric<br />

(ponderosa pine-pinyon-juniper-oak) <strong>for</strong>ests of<br />

<strong>the</strong> Sacramento Mountains <strong>for</strong> two reasons.<br />

First, all five prey species common to this owl<br />

occur in mesic <strong>for</strong>est, whereas only four occur in<br />

xeric <strong>for</strong>ests, where long-tailed voles are absent.<br />

This vole can provide an average mass of 32 g<br />

(1.1 oz) to an owl and it is <strong>the</strong> second most<br />

abundant species occurring in <strong>the</strong> mesic <strong>for</strong>ests.<br />

Second, deer mice dwelling in mesic <strong>for</strong>est can<br />

attain great summer densities during certain<br />

years. This same pattern has not been observed<br />

in <strong>the</strong> seemingly more stable xeric <strong>for</strong>ests. Prey<br />

composition and abundance in ponderosa pine-<br />

Gambel oak <strong>for</strong>ests of nor<strong>the</strong>rn Arizona are<br />

similar to <strong>the</strong> xeric <strong>for</strong>ests of <strong>the</strong> Sacramento<br />

Mountains. In both areas, deer mice are ubiquitous<br />

and <strong>the</strong> <strong>Mexican</strong> woodrat and brush mouse<br />

are patchy in distribution and abundance. In<br />

contrast, <strong>Mexican</strong> voles are apparently less<br />

abundant in pine-oak <strong>for</strong>ests of nor<strong>the</strong>rn Arizona<br />

compared to xeric <strong>for</strong>ests in <strong>the</strong> Sacramento<br />

Mountains. Whe<strong>the</strong>r low densities of voles in<br />

this <strong>for</strong>est type are natural or <strong>the</strong> result of past<br />

management activities is unknown. However,<br />

decreases in herbaceous biomass resulting from<br />

Volume II/Chapter 5<br />

44<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

unnaturally dense <strong>for</strong>est conditions may explain<br />

low numbers of voles in <strong>the</strong>se <strong>for</strong>ests.<br />

Arranging <strong>the</strong> four vegetation communities<br />

according to summer prey biomass indicates that<br />

meadows > mixed-conifer <strong>for</strong>est > ponderosa<br />

pine-pinyon-juniper-oak woodlands > ponderosa<br />

pine-Gambel oak <strong>for</strong>est. When considering<br />

o<strong>the</strong>r factors that influence <strong>the</strong> availability of<br />

prey, mixed-conifer <strong>for</strong>ests likely provide <strong>the</strong><br />

greatest amount of food during summer periods.<br />

Rearranging <strong>the</strong> same communities according to<br />

winter prey biomass indicates that meadows ><br />

ponderosa pine-pinyon-juniper-oak woodlands ><br />

ponderosa pine-Gambel oak <strong>for</strong>est > mixedconifer<br />

<strong>for</strong>est. Accounting <strong>for</strong> <strong>the</strong> availability of<br />

prey, woodlands with a mixture of ponderosa<br />

pine, pinyon-juniper, and oaks provide more<br />

prey to owls during winter months than <strong>the</strong><br />

o<strong>the</strong>r three communities.<br />

Prey biomass is usually greater in all communities<br />

during summer periods when owls raise<br />

<strong>the</strong>ir young. However, temporal peaks of prey<br />

cycles are not correlated among vegetation<br />

communities. That is, when prey are abundant<br />

in mixed-conifer <strong>for</strong>est one year and low a<br />

subsequent year, an opposite pattern may occur<br />

in a different vegetation community. The<br />

asynchrony of abundance among prey species,<br />

vegetation communities, and time may provide a<br />

buffer against <strong>the</strong> effects of extreme oscillation in<br />

prey cycles. This implies <strong>the</strong> importance of<br />

maintaining a mixture of vegetation communities<br />

to ensure a diverse and abundant prey base<br />

within <strong>the</strong> owls <strong>for</strong>aging range.<br />

An important concept exemplified by our<br />

analyses is that <strong>the</strong> habitat of each prey species is<br />

unique. This finding clearly indicates a need <strong>for</strong><br />

providing a variety of conditions which are used<br />

by <strong>the</strong> different species of prey. For example, in<br />

<strong>the</strong> ponderosa pine-Gambel oak <strong>for</strong>ests of<br />

nor<strong>the</strong>rn Arizona, deer mouse abundance shows<br />

little variation according to <strong>for</strong>est structure and<br />

composition whereas <strong>Mexican</strong> woodrat and<br />

brush mouse abundance are strongly correlated<br />

to understory characteristics, specifically log<br />

volume and shrub cover. Fur<strong>the</strong>r, Gambel oak<br />

density is greater within habitats of <strong>the</strong> woodrat<br />

and brush mouse than occurs randomly in <strong>the</strong><br />

<strong>for</strong>est. These habitat components are rarely<br />

considered in planning <strong>for</strong>est management


activities but should be to provide appropriate<br />

habitat conditions <strong>for</strong> <strong>the</strong>se prey species. Obviously,<br />

conserving habitat <strong>for</strong> a diversity of prey<br />

may help buffer against population fluctuations<br />

of individual prey species and provide a less<br />

stochastic food resource <strong>for</strong> <strong>the</strong> owl.<br />

The consequences of three common disturbances<br />

(fire, timber/fuelwood harvest, and<br />

grazing) on <strong>the</strong> owl’s prey and habitat depends<br />

on many factors. Often ecological tradeoffs<br />

result, making exact predictions difficult. Some<br />

prey species may increase, while o<strong>the</strong>rs decrease<br />

<strong>for</strong> a given disturbance. More detailed predictions<br />

about <strong>the</strong> influences of <strong>the</strong>se disturbances<br />

must await more specific research. In general,<br />

management practices that lead to discernible<br />

reductions in total prey biomass or diversity over<br />

large areas will not promote owl recovery.<br />

Volume II/Chapter 5<br />

LITERATURE LITERATURE CITED CITED<br />

CITED<br />

Armstrong, D. M. 1972. Distribution of mammals<br />

in Colorado. Univ. Kansas Mus.<br />

Nat. Hist. Mongr. 3:1-415.<br />

——.1977. Ecological distribution of small<br />

mammals in <strong>the</strong> Upper Williams Fork<br />

Basin, Grand County, Colorado. Southwest.<br />

Nat. 22:289-304.<br />

——.1979. Ecological distribution of rodents in<br />

Canyonlands National Park, Utah. Great<br />

Basin Nat. 39:199-205.<br />

Bailey, V. 1931. Mammals of New Mexico. N.<br />

Amer. Fauna 53:1-412.<br />

Barrows, C. W. 1987. Diet shifts in breeding and<br />

nonbreeding spotted owls. J. Raptor.<br />

Res. 21:95-97.<br />

Campbell, R. E., M. B. Baker, Jr., P. F. Ffolliott,<br />

R. R. Larson, and C. C. Avery. 1977.<br />

Wildfire effects on a ponderosa pine<br />

ecosystem: an Arizona case study. USDA<br />

For. Serv. Res. Pap. RM-191. 12pp.<br />

Carey, A. B. 1985. A summary of <strong>the</strong> scientific<br />

basis <strong>for</strong> spotted owl management. Pages<br />

50-54 in R. J. Gutiérrez, and A. B.<br />

Carey, eds. Ecology and management of<br />

<strong>the</strong> spotted owl in <strong>the</strong> Pacific Northwest.<br />

USDA For. Serv. Gen. Tech. Rep. PNW-<br />

185. 118pp.<br />

Conley, W., and J. D. Nichols. 1978. The use of<br />

models in small mammal population<br />

45<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

studies. Pages 14-35 in D. P. Synder, ed.<br />

Populations of small mammals under<br />

natural conditions. Spec. Publ. No. 5,<br />

Pymatunigh Laboratory of Ecology,<br />

Univ. of Pittsburg, Penn.<br />

Cornely, J. E., and R. J. Baker. 1986. Neotoma<br />

mexicana. Mammal. Species 262:1-7.<br />

——., D. J. Schmidly, H. H. Genoways, and R.<br />

J. Baker. 1981. Mice of <strong>the</strong> Genus<br />

Peromyscus in Guadalupe National Park,<br />

Texas. Occas. Papers Mus. Texas Tech<br />

Univ. 74:1-35.<br />

Craighead, J. J., and F. C. Craighead. 1956.<br />

Hawks, owls and wildlife. Stackpole,<br />

Harrisburg, Penn. 443pp.<br />

DeRosier, S., and J. P. Ward, Jr. 1994. Protocols<br />

<strong>for</strong> analyzing pellets regurgitated by<br />

<strong>Mexican</strong> spotted owls. Unpubl. Rep.,<br />

USDA Forest Service, Rocky Mountain<br />

Research Station, Fort Collins, Colorado.<br />

13pp.<br />

Draper, N. R., and H. Smith. 1981. Applied<br />

regression analysis. Second ed. John<br />

Wiley and Sons, New York, N.Y. 709pp.<br />

Duncan, R. B., and R. Sydner. 1990. Bats in<br />

spotted owl pellets in sou<strong>the</strong>rn Arizona.<br />

Great Basin Nat. 50:197-200.<br />

Durrant, S. D. 1952. Mammals of Utah, taxonomy<br />

and distribution. Univ. Kansas<br />

Publ., Mus. Nat. Hist. 6:1-549.<br />

Errington, P. L. 1930. The pellet analysis<br />

method of raptor food habits study.<br />

Condor 32:292-296.<br />

Findley, J. S., and C. J. Jones. 1982. Distribution<br />

and variation of voles of <strong>the</strong> genus<br />

Microtus in New Mexico and adjacent<br />

areas. J. Mamm. 43:154-166.<br />

——., A. H. Harris, D. E. Wilson, and C. Jones.<br />

1975. Mammals of New Mexico. Univ.<br />

of New Mexico Press, Albuquerque,<br />

N.M. 360pp.<br />

Finerty, J. P. 1980. The population ecology of<br />

small mammals: ma<strong>the</strong>matical <strong>the</strong>ory<br />

and biological fact. Yale Univ. Press, New<br />

Haven, Conn. 234pp.<br />

Finley, R. B., Jr. 1958. The wood rats of Colorado:<br />

distribution and ecology. Univ.<br />

Kansas Mus. Nat. Hist. Pub. 10:213-<br />

552.


——., J. R. Choate, and D. F. Hoffmeister.<br />

1986. Distributions and habitats of voles<br />

in sou<strong>the</strong>astern Colorado and nor<strong>the</strong>astern<br />

New Mexico. Southwest. Nat.<br />

31:263-266.<br />

Forsman, E. D. 1983. Methods and materials <strong>for</strong><br />

locating and studying spotted owls. U.S.<br />

For. Serv. Gen. Tech. Rep. PNW-162.<br />

8pp.<br />

——., C. Meslow, and H. Wight. 1984. Distribution<br />

and biology of <strong>the</strong> spotted owl in<br />

Oregon. Wildl. Monogr. 87. 64pp.<br />

Fox, B. J. 1990. Changes in <strong>the</strong> structure of<br />

mammal communities over successional<br />

time scales. Oikos 59:321-329.<br />

Freeman, M. F., and J. W. Tukey. 1950. Trans<strong>for</strong>mations<br />

related to <strong>the</strong> angular and <strong>the</strong><br />

square root. Ann. Math Statist. 21:607-<br />

611.<br />

Frey, J. A., and C. T. LaRue. 1993. Notes on <strong>the</strong><br />

distribution of <strong>the</strong> mogollon vole<br />

(Microtus mogollonensis) in New Mexico<br />

and Arizona. Southwest. Nat. 38:176-<br />

178.<br />

Ganey, J.L., and J.L. Dick 1995. Habitat relationships<br />

of <strong>the</strong> <strong>Mexican</strong> spotted owl.<br />

Chapter 4 (42 pp.) in <strong>Recovery</strong> plan <strong>for</strong><br />

<strong>the</strong> <strong>Mexican</strong> spotted owl. Vol. II. USDI<br />

Fish and Wildl. Serv., Albuquerque,<br />

NM.<br />

——. 1992. Food habits of <strong>Mexican</strong> spotted<br />

owls in Arizona. Wilson Bull. 104:321-<br />

326.<br />

Getz, L. L. 1985. Habitats. Pages 287-309 in R.<br />

H. Tamarin. ed.. Biology of new world<br />

Microtus. Amer. Soc. Mammalogists<br />

Spec. Publ. No. 8. 839pp.<br />

Glading, B. D., F. Tillotson, and D. M. Selleck.<br />

1943. Raptor pellets as indicators of food<br />

habits. Calif. Fish and Game 29:92-121.<br />

Goodman, L. A. 1960. On <strong>the</strong> exact variance of<br />

products. J. Amer. Stat. Assoc. 55:708-<br />

713.<br />

Goodwin, J. G., Jr., and C. R. Hunger<strong>for</strong>d.<br />

1979. Rodent population densities and<br />

food habitats in Arizona ponderosa pine<br />

<strong>for</strong>ests. USDA For. Serv. Res. Pap. RM-<br />

214. 12pp.<br />

Grant, W. E., E. C. Birney, N. R. French, and<br />

D. M. Swift. 1982. Structure and pro-<br />

Volume II/Chapter 5<br />

46<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

ductivity of grassland small mammal<br />

communities related to grazing induced<br />

changes in vegetative cover. J. Mammal.<br />

63:248-260.<br />

Hall, E. R., and K. R. Kelson. 1959. The mammals<br />

of North America. Ronald Press,<br />

Co., N.Y. 2 vols. 1083pp.<br />

Hanley, T. A., and J. L. Page. 1982. Differential<br />

effects of livestock use on habitat structure<br />

and rodent populations in Great<br />

Basin communities. Calif. Fish and<br />

Game 68:160-174.<br />

Hirons, G. J. M. 1985. The importance of body<br />

reserves <strong>for</strong> successful reproduction in<br />

<strong>the</strong> Tawny owl (<strong>Strix</strong> aluco). J. Zool.<br />

Lond. (B) 1:1-20.<br />

Hoffmeister, D. F. 1986. Mammals of Arizona.<br />

Univ. of Arizona Press, Tucson, Ariz.<br />

602pp.<br />

Johnsgard, P. A. 1988. North American owls:<br />

biology and natural history. Smithsonian<br />

Inst. Press, Washington, D.C. 295pp.<br />

Johnson, D. W., and D. M. Armstrong. 1987.<br />

Peromyscus crinitus. Mammal. Species<br />

287:1-8.<br />

Kauffman, J. B., and W. C. Krueger. 1984.<br />

Livestock impacts on riparian ecosystems<br />

and streamside management implications,<br />

a review. J. Range Manage.<br />

37:430-438.<br />

——., ——., and M. Vavra. 1983. Effects of late<br />

season cattle grazing on riparian plant<br />

communities. J. Range Manage. 36:685-<br />

691.<br />

Kaufman, G. A., D. W. Kaufman, and E. J.<br />

Finck. 1988. Influence of fire and<br />

topography on habitat selection by<br />

Peromyscus maniculatus and<br />

Reithrodontomys megalotis in ungrazed<br />

tall-grass prairie. J. Mammal. 66:342-<br />

352.<br />

Kertell, K. 1977. The spotted owl at Zion<br />

National Park, Utah. West. Birds 8:147-<br />

150.<br />

Kirkland, G. L., Jr. 1990. Patterns of initial<br />

small mammal community change after<br />

clearcutting of temperate North American<br />

<strong>for</strong>ests. Oikos 59:313-320.<br />

Korpimäcki, E., and K. Norrdahl. 1991. Numerical<br />

and functional responses of


kestrels, short-eared owls, and long-eared<br />

owls to vole densities. Ecology 72:814-<br />

826.<br />

Lundberg, A. 1976. Breeding success and prey<br />

availability in a Ural owl (<strong>Strix</strong> uralensis<br />

Pall.) population in central Sweden.<br />

Zoon 4:65-72.<br />

Marti, C. D. 1987. Raptor food habits studies.<br />

Pages 67-80 in B. A. Giron Pendelton,<br />

ed. Raptor Management Techniques<br />

Manual. National Wildlife Institute,<br />

Washington, D.C. 394pp.<br />

McNaughton, S. J. 1993. Grasses and grazers,<br />

science and management. Ecol. Applic.<br />

3:17-20.<br />

Medin, D. E., and W. P. Clary. 1990. Bird and<br />

small mammal populations in a grazed<br />

and ungrazed riparian habitat in Idaho.<br />

USDA For. Serv. Res. Pap. INT-425.<br />

8pp.<br />

Milchunas, D. G., and W. K. Lauenroth. 1993.<br />

Quantitative effects of grazing on vegetation<br />

and soils over a global range of<br />

environments. Ecol. Monogr. 63:327-<br />

366.<br />

Newton, I. 1979. Population ecology of raptors.<br />

Vermillion, S.D. 399pp.<br />

Orodho, A. B., M. J. Trlica, and C. D. Bonham.<br />

1990. Long-term heavy-grazing effects<br />

on soil and vegetation in <strong>the</strong> four corners<br />

region. Southwest. Nat. 35:9-14.<br />

Otis, D. L., K. P. Burnham, G. C. White, and<br />

D. R. Anderson. 1978. Statistical inference<br />

from capture data on closed animal<br />

populations. Wildl. Monogr. 62. 135pp.<br />

Peterson, S. K., G. A. Kaufman, and D. W.<br />

Kaufman. 1985. Habitat selection by<br />

small mammals of <strong>the</strong> tall-grass prairie:<br />

experimental patch choice. Prairie Nat.<br />

17:65-70.<br />

Putnam, R. J. 1986. Grazing in temperate<br />

ecosystems, large herbivores and <strong>the</strong><br />

ecology of <strong>the</strong> New Forest. Timber Press,<br />

Portland, Oreg. 210pp.<br />

Reichenbacher, F. W., and R. B. Duncan. 1992.<br />

Diet of <strong>Mexican</strong> spotted owls on National<br />

Forest system lands in Arizona and<br />

New Mexico. Unpubl. Rep., USDA For.<br />

Serv., Southwestern Regional Office,<br />

Volume II/Chapter 5<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

Albuquerque, N.M. 77pp.<br />

Reynolds, R. T., R. T. Graham, M. H. Reiser, R.<br />

L. Bassett, P. L. Kennedy, D. A. Boyce,<br />

Jr., G. Goodwin, R. Smith, and E. L.<br />

Fisher. 1992. Management recommendations<br />

<strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn goshawk in <strong>the</strong><br />

southwestern United States. USDA For.<br />

Serv. Gen. Tech. Rep. RM-217. 90pp.<br />

Reynolds, T. D. 1980. Effects of some different<br />

land management practices on small<br />

mammal populations. J. Mammal.<br />

61:558-561.<br />

Ribble, D. O., and F. B. Samson. 1987. Microhabitat<br />

associations of small mammals in<br />

sou<strong>the</strong>astern Colorado, with special<br />

emphasis on Peromyscus (Rodentia).<br />

Southwest. Nat. 32:291-303.<br />

Rinkevich, S. E. 1991. Distribution and habitat<br />

characteristics of <strong>Mexican</strong> spotted owls<br />

in Zion National Park, Utah. M.S.<br />

Thesis. Humboldt State Univ., Arcata,<br />

Calif. 62pp.<br />

Schulz, T. T., and W. C. Leininger. 1990. Differences<br />

in riparian vegetation structure<br />

between grazed areas and exclosures. J.<br />

Range Management 43:294-298.<br />

——. 1991. Nongame wildlife communities in<br />

grazed and ungrazed montane riparian<br />

sites. Great Basin Nat. 51:286-292.<br />

Sou<strong>the</strong>rn, H. N. 1970. The natural control of a<br />

population of Tawny owls (<strong>Strix</strong> aluco). J.<br />

Zool. 162:197-285.<br />

Svoboda, P. L., D. K. Tolliver, and J. R. Choate.<br />

1988. Peromyscus boylii in <strong>the</strong> San Luis<br />

Valley, Colorado. Southwest. Nat.<br />

33:239-240.<br />

Szaro, R. C. 1991. Wildlife communities of<br />

southwestern riparian ecosystems. Pages<br />

174-201 in Rodiek, J. E., and E. G.<br />

Bolen eds. Wildlife and habitats in<br />

managed landscapes. Island Press, Washington,<br />

D.C. 219pp.<br />

Tarango, L. A. 1994. <strong>Mexican</strong> spotted owl distribution<br />

and habitat characterizations in southwestern<br />

Chihuahua, Mexico. M.S. Thesis,<br />

New Mexico State Univ., Las Cruces, N.M.<br />

Thomas, J. W., E. D. Forsman, J. B. Lint, E. C.<br />

Meslow, B. R. Noon, and J. Verner. 1990.<br />

A conservation strategy <strong>for</strong> <strong>the</strong> spotted owl.<br />

U.S. Gov. Print. Off., Washington, D.C.<br />

427pp.<br />

47


Thrailkill, J., and M. A. Bias. 1989. Diets of breeding<br />

and nonbreeding Cali<strong>for</strong>nia spotted<br />

owls. J. Raptor Res. 23:39-41.<br />

USDA Forest Service. 1994. Biological assessment<br />

and evaluation: livestock grazing. Unpubl.<br />

Rep., USDA For. Serv., Southwest Region,<br />

Albuquerque, N.M. 12pp.<br />

USDI Fish and Wildlife Service. 1995. <strong>Recovery</strong><br />

<strong>Plan</strong> <strong>for</strong> <strong>the</strong> <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong>. Vol. 1.<br />

USDI Fish and Wildl. Serv., Albuquerque,<br />

N.M. 178pp.<br />

Utah <strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> Technical. Team. 1994.<br />

Suggestions <strong>for</strong> management of <strong>Mexican</strong><br />

spotted owls in Utah. Utah Div. of Wildl.<br />

Resour., Salt Lake City, Utah. 103pp.<br />

Vallentine, J. F. 1990. Grazing management. Academic<br />

Press, Inc., San Diego, Calif. 533pp.<br />

Verner, J., K. S. McKelvey, B. R. Noon., R. J.<br />

Gutiérrez, G. I. Gould, Jr., and T. W. Beck,<br />

eds. 1992. The Cali<strong>for</strong>nia spotted owl: a<br />

technical assessment of its current status.<br />

USDA For. Serv. Gen. Tech. Rep. PSW-133.<br />

285pp.<br />

——., M. L. Morrison, and C. J. Ralph. 1986.<br />

Wildlife 2000: modeling habitats of terrestrial<br />

vertebrates. Univ. Wisc. Press, Madison,<br />

Wisc. 470pp.<br />

Wagner, P. W., C. D. Marti, and T. C. Boner. 1982.<br />

Food of <strong>the</strong> spotted owl in Utah. J. Raptor<br />

Res. 16:27-28.<br />

Ward, J. P., Jr. 1990. <strong>Spotted</strong> owl reproduction, diet<br />

and prey abundance in northwest Cali<strong>for</strong>nia.<br />

M.S. Thesis, Humboldt State Univ., Arcata,<br />

Calif. 70pp.<br />

Wendland, V. 1984. The influence of prey fluctuations<br />

on <strong>the</strong> breeding success of <strong>the</strong> Tawny<br />

owl (<strong>Strix</strong> aluco). Ibis 126:285-295.<br />

White, G. C., D. R. Anderson, K. P. Burnham, and<br />

D. L. Otis. 1982. Capture-recapture and<br />

removal methods <strong>for</strong> sampling closed<br />

populations. USDE Los Alamos National<br />

Lab. Rep. LA-8787-NERP. Los Alamos,<br />

N.M. 235pp.<br />

Wilson, D. E. 1968. Ecological distribution of <strong>the</strong><br />

genus Peromyscus. Southwest. Nat. 13:267-<br />

274.<br />

Volume II/Chapter 5<br />

48<br />

<strong>Mexican</strong> <strong>Spotted</strong> <strong>Owl</strong> <strong>Recovery</strong> <strong>Plan</strong><br />

APPENDIX APPENDIX 5 55a<br />

5<br />

Sources <strong>for</strong> estimates of prey mass used in calculating<br />

percent prey biomass in <strong>Mexican</strong> spotted owl diets.<br />

Anderson, S. 1972. Mammals of Chihuahua,<br />

taxonomy and distribution. Bull. Amer.<br />

Museum Nat. Hist. 148:149-410.<br />

Armstrong, D. M. 1972. Distribution of mammals<br />

in Colorado. Mongr. Univ. Kansas Mus.<br />

Nat. Hist. 3:1-415.<br />

Berna, H. J. 1990. Seven bat species from <strong>the</strong> Kaiba<br />

Plateau, Arizona, with a record of Euderma<br />

maculatum. Southwest. Nat. 35:354-356.<br />

Block, W. M., and J. L. Ganey. Unpublished small<br />

mammal masses collected in Nor<strong>the</strong>rn<br />

Arizona (1990-1992). USDA For. Serv.,<br />

Rocky Mtn. For. Range Exp. Sta., Flagstaff,<br />

Arizona.<br />

Bogan, M., and C. Ramotnik. Unpublished small<br />

mammal masses collected in Utah (1990-<br />

1992). USDI Fish and Wildl. Serv., Mus. of<br />

Southwest Biol., Univ. of New Mexico,<br />

Albuquerque, N.M.<br />

Burt, W. H., and R. P. Grossenheider. 1976. A field<br />

guide to <strong>the</strong> mammals. Hougton Mifflin<br />

Co., Boston, Mass. 289 pp.<br />

Chapman, J. A., and G. A. Feldhammer. 1982. Wild<br />

mammals of North America. Johns<br />

Hopkins Univ. Press, Baltimore, Maryl.<br />

1147 pp.<br />

Dunning, J. B. ed. 1993. CRC handbook of avian<br />

masses. CRC Press, Boca Raton, Flor. 371<br />

pp.<br />

Ganey, J. L. Unpublished masses used to determine<br />

prey biomass in diet of <strong>Mexican</strong> spotted owls<br />

presented in 1992 Wilson Bull. publication.<br />

Johnsgard, P. A. 1988. North American owls, biology<br />

and natural history. Smithsonian Institution<br />

Press, Wash. D. C. 285 pp.<br />

Marti, C. D. 1974. Feeding ecology of four sympatric<br />

owls. Condor 76:45-61.<br />

Severson, K. E., and B. J. Hayward. 1980. Rodent<br />

weights in modified pinyon-juniper woodlands<br />

of southwestern New Mexico. Great<br />

Basin Nat. 48:554-557.<br />

Steenhof, K. 1983. Prey weights <strong>for</strong> computing<br />

percent biomass in raptor diets. Raptor Res.<br />

17:15-27.<br />

Ward, J. P., Jr., S. DeRosier, and W. M. Block.<br />

Unpublished masses of small mammals<br />

collected in <strong>the</strong> Sacramento Mountains,<br />

New Mexico (1991-1993). Rocky Mtn. For.<br />

Range Exp. Sta., Ft. Collins, Colo.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!